Ocean Fertilization: Science, Policy, and Commerce Please share

advertisement
Ocean Fertilization: Science, Policy, and Commerce
The MIT Faculty has made this article openly available. Please share
how this access benefits you. Your story matters.
Citation
Strong, Aaron L., John J. Cullen, and Sallie W. Chisholm.
"Ocean Fertilization: Science, Policy, and Commerce."
Oceanography 22.3 (2009): 236–261.
As Published
http://www.tos.org/oceanography/issues/issue_archive/issue_pdf
s/22_3/22-3_strong.pdf
Publisher
The Oceanography Society
Version
Final published version
Accessed
Wed May 25 22:06:22 EDT 2016
Citable Link
http://hdl.handle.net/1721.1/77572
Terms of Use
Article is made available in accordance with the publisher's policy
and may be subject to US copyright law. Please refer to the
publisher's site for terms of use.
Detailed Terms
Ocean
Fertilization
Science, Policy, and Commerce
B y A a r on L . S t r on g , J o h n J . C u ll e n ,
an d S all i e W. C h i s h ol m
ABSTR ACT. Over the past 20 years there has been growing interest in the
concept of fertilizing the ocean with iron to abate global warming. This interest
was catalyzed by basic scientific experiments showing that iron limits primary
production in certain regions of the ocean. The approach—considered a form of
“geoengineering”—is to induce phytoplankton blooms through iron addition, with
the goal of producing organic particles that sink to the deep ocean, sequestering
carbon from the atmosphere. With the controversy surrounding the most recent
scientific iron fertilization experiment in the Southern Ocean (LOHAFEX) and
the ongoing discussion about restrictions on large-scale iron fertilization activities
by the London Convention, the debate about the potential use of iron fertilization
for geoengineering has never been more public or more pronounced. To help
inform this debate, we present a synoptic view of the two-decade history of iron
fertilization, from scientific experiments to commercial enterprises designed to
trade credits for ocean fertilization on a developing carbon market. Throughout
these two decades there has been a repeated cycle: Scientific experiments are
followed by media and commercial interest and this triggers calls for caution and
the need for more experiments. Over the years, some scientists have repeatedly
pointed out that the idea is both unproven and potentially ecologically disruptive,
and models have consistently shown that at the limit, the approach could not
substantially change the trajectory of global warming. Yet, interest and investment
in ocean fertilization as a climate mitigation strategy have only grown and
intensified, fueling media reports that have misconstrued scientific results, and
conflated scientific experimentation with geoengineering. We suggest that it is
time to break this two-decade cycle, and argue that we know enough about ocean
fertilization to say that it should not be considered further as a means to mitigate
climate change. But, ocean fertilization research should not be halted: if used
appropriately and applied to testable hypotheses, it is a powerful research tool for
understanding the responses of ocean ecosystems in the context of climate change.
236
Oceanography
Vol.22, No.3
This article has been published in Oceanography, Volume 22, Number 3, a quarterly journal of The Oceanography Society. © 2009 by The Oceanography Society. All rights reserved. Permission is granted to copy this article for use in teaching and research. Republication, systemmatic reproduction,
or collective redistirbution of any portion of this article by photocopy machine, reposting, or other means is permitted only with the approval of The Oceanography Society. Send all correspondence to: info@tos.org or Th e Oceanography Society, PO Box 1931, Rockville, MD 20849-1931, USA.
S c i e nc e an d P ol i c y F e at u r e
INTRODUCTION
Over the 20 years since oceanographer
John Martin of Moss Landing Marine
Laboratories quipped, “Give me half
a tanker of iron, and I’ll give you an
ice age,” fertilization of the ocean with
iron has drawn increasing attention as
a potential geoengineering strategy for
carbon sequestration. As interest in the
idea has increased, so has the controversy surrounding it. In January 2009,
major news services broadcast the
gripping story of the suspension of
LOHAFEX1, an iron fertilization
experiment in the Southern Ocean.
The research vessel Polarstern was
midway between South Africa and South
America when the German Research
Ministry put a halt to the experiment.
The story was featured in Wired magazine (Keim, 2009) and on Reuters (Szabo,
2009) and the BBC (Morgan, 2009). A
blog headline covering the controversy
posed the question: “LOHAFEX—If
you mean well, are you allowed to screw
up the oceans?” (Campbell, 2009). The
experiment drew immediate and intense
commentary from environmental
nongovernmental organizations (NGOs),
one of which cast it as a violation of the
2008 United Nations (UN) Convention
on Biological Diversity’s moratorium on
iron fertilization activities (ETC Group
News Release, 2009). The international
press, including news articles in Science,
consistently referred to LOHAFEX as
a “geoengineering project,” an experiment designed to test the potential of
ocean iron fertilization to change global
climate (Kintisch, 2009).
1
The LOHAFEX scientists, however,
defended their experiment as purely
scientific and consistent with relevant
UN regulations. And, the Director of the
Alfred Wegener Institute (AWI), which
sponsored LOHAFEX, defended the
scientific validity of the research, stating
that they “neither plan to nor want to
seeking to carry out large-scale ocean
fertilization activities to sell carbon-offset
credits in a carbon trading market. These
ventures have been a cause of concern
for some scientists and several environmental NGOs, who have argued that
claims of significant carbon sequestration
are unsupported by scientific evidence,
“
…the histories of the scientific and
commercial interests in ocean iron
fertilization (OIF) are intimately
connected—co-evolving and
transforming over time.
smooth the way for a commercial use of
iron fertilization with our expedition,”
and that they “oppose iron fertilization
with the aim to reduce CO2 to regulate
the climate” (AWI News, 2009). After
the German Research Ministry received
several independent environmental
assessments of LOHAFEX and an impact
statement from the Indo-German
research crew on Polarstern, the green
light was given (AWI Press Release,
2009a): scientists commenced fertilizing
the Southern Ocean with 10 tonnes of
iron sulfate on January 27, 2009.
At the core of the controversy over
LOHAFEX was the idea of using iron
fertilization to mitigate global climate
change by sequestering carbon dioxide
in the deep ocean. Over the last two
decades, this idea has attracted the
interest of several commercial ventures
”
and that large-scale iron fertilization
will, by design, profoundly alter marine
ecosystems (Chisholm et al., 2001;
Gnanadesikan et al., 2003; Cullen and
Boyd, 2008; Denman, 2008; ETC Group
News Release, 2009; World Wildlife Fund
International, 2009). These concerns have
helped spur recent UN resolutions, which
were intended to restrict iron fertilization
activities to small-scale scientific research
(UN CBD, 2008; London Convention
Meeting Report, 2008).
As we describe below, the histories of
the scientific and commercial interests
in ocean iron fertilization (OIF) are
intimately connected—co-evolving and
transforming over time. In studying
this history, it becomes apparent that
despite the lack of experimental results
indicating that OIF would be effective for significant climate mitigation,
“LOHA” is Hindi for “iron,” FEX stands for “FErtilization eXperiment.”
Oceanography
September 2009
237
commercial interests have continued to
pursue and advance the idea, fueling a
cycle of media interest, followed by calls
for caution, and then proposals for more
research requiring longer and larger
experiments. Here we take a synoptic
view of the two-decade history of this
cycle (Figure 1) to better understand
how the scientific and commercial
interests have become intertwined,
conflated, and confused. We have much
to learn from this history, as proposals
for research on other forms of geoengineering are beginning to emerge
(Latham et al., 2008; Rasch et al., 2008).
Figure 1. Timeline of ocean iron
fertilization experiments (OIF),
private sector interest in them,
and policy developments.
HISTORY OF PUBLICSECTOR SCIENTIFIC O CE AN
FERTILIZATION E XPERIMENTS
The “Iron Hypothesis”
While the idea that iron limitation might
control productivity in certain areas of
the ocean has been around since the
mid-1920s (Hart, 1934) if not before
(de Baar, 1994), the contemporary
history of ocean iron fertilization and the
“iron hypothesis” is attributed to John
Martin. During a 1988 lecture at the
Woods Hole Oceanographic Institution,
Aaron L. Strong is Research Assistant,
Department of Civil and Environmental
Engineering, Massachusetts Institute
of Technology, Cambridge, MA, USA.
John J. Cullen (john.cullen@dal.ca)
is Killam Chair of Ocean Studies,
Department of Oceanography, Dalhousie
University, Halifax, Nova Scotia, Canada.
Sallie W. Chisholm (chisholm@mit.edu) is
Martin Professor of Environmental Studies,
Department of Civil and Environmental
Engineering, Department of Biology,
Massachusetts Institute of Technology,
Cambridge, MA, USA.
238
Oceanography
Vol.22, No.3
Martin famously quipped, “Give me
half a tanker of iron, and I’ll give you
an ice age” (Martin, 1990a). Referring
to shipboard experiments on samples
of seawater, and ice core data showing a
relationship between atmospheric iron
dust deposition and atmospheric CO2
concentrations over the past 160,000
years (Figure 2), Martin developed a
two-part hypothesis. First, he argued
that the “high nutrient, low chlorophyll”
(HNLC) regions of the ocean could be
explained by iron limitation (i.e., surface
nitrate and phosphate concentrations
were not depleted because additional
phytoplankton growth was limited by
iron). Second, he argued that if iron
did indeed control the productivity of
HNLC waters, and thus the transport of
organic carbon to the deep sea via the
so-called biological pump (Figure 3), it
could explain the observed relationship
between atmospheric iron dust deposition and atmospheric CO2 during the
last glacial maximum (Martin, 1990b).
He also commented that this “paleoiron” hypothesis could be important,
because intentional oceanic iron fertilization could prove an effective method
of drawing down atmospheric CO2
“should the need arise” (Martin et al.,
1990; see Box 1).
Using ultra-clean experimental
approaches with seawater incubated on
the deck of a ship, Martin’s team successfully demonstrated that iron addition
could stimulate phytoplankton growth
in bottled samples collected from the
HNLC Gulf of Alaska (Martin et al.,
1989) and the HNLC Southern Ocean
(Martin et al., 1990). Wide acceptance of
the iron hypothesis would require tests
on open waters, however. Though, sadly,
Martin did not live to see it, in 1993 his
Figure 2. Relationship between Fe and CO2 concentrations from the
Vostok ice cores. After Martin (1990b)
colleagues carried out the first openocean mesoscale (i.e., more than several
kilometers on a side) iron fertilization
experiment, “IronEx I,” in equatorial
Pacific HNLC waters (Figure 4). The
results demonstrated a phytoplankton
bloom in response to iron addition,
but they were confounded when the
fertilized patch was subducted under
low-density water (Martin et al., 1994).
Thus, several hypotheses remained
untested (Cullen, 1995).
Armed with this experience, a second
experiment, IronEx II, was organized
in May 1995—again in the equatorial
Pacific. This time, a 72-km2 patch was
fertilized serially three times over the
course of one week. This experiment
demonstrated a strong phytoplankton
response to iron addition in these
HNLC waters, prompting the authors
to conclude that “it is now time to
regard the ‘iron hypothesis’ as the
‘iron theory’” (Coale et al., 1996). The
authors suggested that the logical next
step was to conduct an experiment in
the Southern Ocean as this is “where
most of the HNLC waters are found and
where paleoclimate coherence between
iron flux and carbon export has been
observed” (Coale et al., 1996).
The Next Round of Experiments:
Carbon Sequestration
Becomes the Hypothesis
Over the five years from 1999 to 2004,
eight more major open-ocean iron fertilization experiments would take place in
the Southern Ocean and subarctic Pacific
HNLC regions (Figure 4; Table 1). These
experiments sought, for the most part,
to track the fate of carbon fixed in ironinduced blooms. As such, they became
increasingly focused on the question of
carbon export from the surface waters,
as this is what is necessary to draw CO2
out of the atmosphere and transport it
into the deep sea. At the same time, as
interests in OIF for carbon sequestration were taking off (see below), the
scientific language began to morph:
increasingly, “carbon export” became
Oceanography
September 2009
239
Figure 3. Carbon dioxide that would
otherwise be in the atmosphere is
stored in the deep sea because the
biological pump puts and keeps it
there. Phytoplankton in the lighted
surface layer take up nutrients
(e.g., nitrate and phosphate) and
grow, converting CO2 to organic
matter that fuels marine food webs.
Some of the organic matter—for
example, senescent phytoplankton,
fecal pellets, and aggregated debris—
sinks to the deep ocean where it
decomposes, releasing CO2 and
nutrients while consuming oxygen.
When the ocean carbon cycle is
roughly in balance, this carbon- and
nutrient-rich deep water does not
reach the surface for decades to
hundreds of years, and when it does,
biological productivity consumes the
CO2 and nutrients and sends C, N,
and P back to deep waters as sinking
organic matter. The amount of CO2
thus stored in the deep sea largely
corresponds to the amount of major
nutrients (N and P) consumed in the
lighted surface layer of the ocean.
When iron limits productivity, N and
P persist where they wouldn’t otherwise; if iron limitation is alleviated,
major nutrients are consumed, more
organic matter is produced, and more
carbon sinks to the deep sea. This extra carbon associated with added iron (either natural or intentional)
could be considered sequestration. But the amount and duration of carbon sequestration depends on how
deep the organic matter sinks before it is decomposed and whether or not iron is still available in excess
when carbon- and nutrient-enriched waters reach the surface again. Figure modified from Chisholm (2000)
“carbon sequestration” (Boyd et al., 2000,
2004; Buesseler et al., 2004), a term
used in the US Department of Energy
(DOE) climate mitigation vocabulary
(Department of Energy, 1999). This
shift in language contributed to blurring
the lines between the basic and applied
science dimensions of OIF.
An analysis of the “post-IronEx”
mesoscale experiments illuminates the
evolution of hypotheses that motivated
them (Table 1; see de Baar et al., 2005,
and Boyd et al., 2007, for reviews). The
240
Oceanography
Vol.22, No.3
new focus on the Southern Ocean began
with the Southern Ocean Iron Release
Experiment (SOIREE) and the European
Iron Enrichment Experiment (EisenEx;
Boyd et al., 2000; Smetacek, 2001),
and placed more emphasis on longerduration experiments tracking particle
export, remineralization, and changes in
zooplankton communities. Both experiments confirmed the hypothesis of iron
limitation of primary production in
the HNLC Southern Ocean. Although
diatom production increased in response
to iron addition in the SOIREE patch,
carbon export did not (Boyd et al., 2000).
EisenEx also demonstrated a diatom
bloom, and measured a larger net atmospheric CO2 drawdown than SOIREE,
but storms interrupted the experiment
and the fate of fixed carbon could not
be tracked (Assmy et al., 2007). A year
later, SEEDS-I (Subarctic Pacific Iron
Experiment for Ecosystem Dynamics
Study) confirmed that productivity
was limited by iron in the HNLC
region of the western subarctic Pacific,
Box 1. The Paleo-Climate Portion of the Iron Hypothesis: Still an Open Question
Recently, some paleoceanographers have called into question the causal link between atmospheric dust deposition and lower atmospheric CO2
(and thus a cooler climate) during the last glacial maximum. Kohfeld et al. (2005) analyzed the role of the biological pump in glacial CO2 drawdown using sediment records, for example. Their data indicate that in large portions of the Southern Ocean, export productivity was actually
lower during the last glacial maximum—exactly during a period of increased dust flux. They thus argued that iron fertilization from dust could
not have been solely responsible for CO2 drawdown (Kohfeld et al., 2005).
Using these productivity data and those of others (Paytan et al., 1996; Anderson et al., 2008) and comparing them to dust flux records from
Winckler et al. (2008), Anderson et al. (2007) presented an argument that there is no correlation in the paleoceanographic data between dust
flux and increased export productivity in the equatorial Pacific or the Southern Ocean. While their data do show a strong anti-correlation
between dust flux and CO2, they argue that there is no evidence that the dust caused the CO2 drawdown. The causality inferred from the
original ice core data (Figure 2) has been a central thread in the argument for OIF for geoengineering, and, at the very least, these recent arguments questioning that causality deserve more attention and research (Anderson et al., 2007). Alternate hypotheses include strong influences
of changing wind patterns on the overturning of carbon-rich southern deep water (Toggweiler et al., 2006).
documenting a floristic shift toward
diatom production. Carbon export was
not measured (Tsuda et al., 2005).
The next set of experiments were
designed to be larger and longer, with
the hope that this would allow better
tracking of the fate of iron-induced
blooms. The Southern Ocean Iron
Experiment (SOFeX), a multi-part,
multi-ship iron fertilization expedition, set out to address hypotheses
about carbon export as influenced by
silicate availability in the context of
iron-induced blooms (Coale et al., 2004).
SOFeX produced the first conclusive
measurement of enhanced particulate
organic carbon (POC) export resulting
from an intentional iron-fertilizationinduced bloom (Buesseler et al.,
2004). Although POC export from the
iron-enriched patch was elevated, the
incremental flux was small compared
to those observed during a natural iron
fertilization event in the same location
in 1998 (Buesseler et al., 2001). The
authors concluded that the observed
carbon flux was small relative to what
would be required by proposed geoengineering plans to sequester carbon using
iron fertilization, but they stressed that
they had not been able to observe the
termination of the iron-induced bloom
(Buesseler et al., 2004).
The Subarctic Ecosystem Response
to Iron Enrichment Study (SERIES)
expedition followed. Conducted over a
month in the subarctic Pacific, a 77-km2
iron-enriched patch was created and
POC export flux monitored (Boyd
et al., 2004). The decline and termination of the iron-induced bloom (caused
by silicate limitation) was observed.
The majority of the carbon fixed in
SERIES was remineralized by bacteria
and zooplankton grazing in the surface
waters. Only a small fraction (8%) of
the fixed carbon sank below the 120-m
permanent pycnocline, significantly
lower than the deep-export rate observed
in natural blooms (Buesseler, 1998).
Further, the iron content (Fe:C) of the
exported material was a thousandfold
higher than that assumed in assessments of the efficiency and cost of OIF
for climate mitigation. Noting increased
attention to the idea of iron fertilization
for geoengineering, Boyd et al. (2004)
argued that “inefficient vertical transfer
of carbon may limit the effectiveness of
iron fertilization as a mitigation strategy.”
The European Iron Fertilization
Experiment (EIFEX), conducted in
2004 and the longest iron fertilization
experiment to date, was also designed to
evaluate the carbon export response and
community shifts in a Southern Ocean
iron-induced bloom (Hoffmann et al.,
2006). Biomass export resulting from
the sinking of an iron-induced bloom
represented the highest ratio of carbon
exported to added iron to date (Jacquet
et al., 2008). At the same time, SEEDS-II,
a second subarctic Pacific experiment,
detected no significant bloom response
to iron enrichment (Tsuda et al., 2007).
Oceanography
September 2009
241
Figure 4. Locations of major artificial iron enrichment experiments, including the pilot demonstrations of GreenSea Venture
and Planktos. Color heat map represents surface nitrate concentrations with warmer colors indicating higher concentrations,
showing three major HNLC regions in the Southern Ocean, the eastern equatorial Pacific, and the subarctic Pacific. Data from
National Virtual Ocean Data System, http://ferret.pmel.noaa.gov/NVODS/; analyzed nitrate data from the World Ocean Atlas 2005
Two other experiments were
conducted in 2004 to test additional
biogeochemical hypotheses using
iron enrichment. The Surface-Ocean
Lower-Atmosphere Studies Air-Sea Gas
Exchange (SAGE) experiment in the
sub-Antarctic Pacific off New Zealand
sought to understand the effects of iron
addition on air-sea gas exchange, particularly dimethylsulfide (DMS). DMS
plays a role in cloud formation, and as
such is thought to play a role in climate
regulation and, potentially, in increasing
Earth’s albedo (Charlson et al., 1987).
In this experiment, scientists fertilized
serially with over 5 tonnes of iron sulfate,
but found only a modest chlorophyll
increase and no increase in either CO2
drawdown or DMS production, which
242
Oceanography
Vol.22, No.3
they believed may have been due to light
limitation (Law et al., 2006). Another
iron-enrichment experiment, FeeP,
performed a combined Fe and phosphate
addition to low nutrient, low chlorophyll (LNLC) waters in the Northeast
Atlantic to test the hypothesis that iron
enrichment of LNLC waters can lead to
a net N import (ultimately supporting
increased productivity) by stimulating
nitrogen fixation. Although rates of
nitrogen fixation increased in response
to enrichment, productivity did not.
Carbon export in response to enrichment was not measured (Rees et al.,
2007; Karl and Letelier, 2008).
Collectively, these experiments taught
us a good deal about the initial phytoplankton community response to iron
enrichment, and confirmed iron limitation of productivity in HNLC regions
around the globe. The results, however,
were highly variable and inconclusive
with regard to carbon sequestration: the
major conclusion to be drawn is that
physical oceanographic, geographic, and
biological variability all influence the
long-term fate of blooms induced by iron
fertilization, significantly constraining
generalizations about iron-induced
carbon sequestration (Boyd et al., 2007).
Modeling
Open ocean iron enrichment experiments over the last 16 years have
confirmed the hypothesis of iron limitation of productivity in HNLC regions
and have provided evidence for highly
Table 1. Summary of ocean iron enrichment experiments conducted between 1993 and 2009.
See also reviews by de Baar et al. (2005) and Boyd et al. (2007).
Experiment
Year
Location
Duration
Magnitude
Rationale/Hypothesis Tested
General Conclusions
IronEx I
Martin et al., 1994
1993
Eastern equatorial
Pacific Ocean
10 days
450 kg Fe
64 km2
• Iron limitation of productivity
in HNLC region
• Iron limits phytoplankton growth
rate, but patch subducted; broader
implications of OIF unclear
IronEx II
Coale et al., 1996
1995
Eastern equatorial
Pacific Ocean
17 days
450 kg Fe
72 km2
• Iron limitation of productivity
in HNLC region
• Iron definitively limits productivity
in equatorial Pacific. Larger bloom
than IronEx I
1999
Southern OceanAustralia; South
of Antarctic Polar
Front (APF)
13 days
1740 kg Fe
50 km2
• Iron limitation of productivity
in Southern Ocean, south of
the Antarctic Polar Front (APF)
• Fate of carbon fixed in bloom
• Iron limits productivity in
Southern Ocean
• No downward carbon transport
observed
• Iron limits productivity in
Southern Ocean
• Fate of the bloom uncertain
• Iron only affected certain species
of phytoplankton
SOIREE
Boyd et al., 2000
EisenEx
Smetacek, 2001;
Assmy et al., 2007
2000
Southern OceanAfrica; in APF zone
21 days
4 tonnes FeSO4
38.5 km2
• Iron limitation of productivity
in Southern Ocean, along APF
• Simulate Fe dust deposition to
test whether Fe dust contributed to lower atmospheric
CO2 concentrations during
glacial periods
SEEDS-I
Tsuda et al., 2005
2001
Subarctic PacificNorthwest
13 days
350 kg Fe
80 km2
• Iron limitation of productivity
in HNLC of subarctic Pacific
• Fate of carbon fixed in bloom
• Iron limits productivity in
subarctic Pacific
• Floristic shift to diatoms
• Downward carbon export minimal
2002
Southern OceanNew Zealand;
north and south
of APF
30 days
N: 1712 kg Fe
225 km2
S: 1260 Kg Fe
225 km2
• Does OIF increase flux of
carbon to deep ocean?
• Silicate influence and
geographic variability of
response
• Increase in POC export flux, but
magnitude is small relative to
natural blooms
2002
Subarctic PacificGulf of Alaska
25 days
490 kg Fe
77 km2
• Fate of carbon fixed in ironinduced bloom
• Efficiency of carbon export to
deep ocean
• Majority of carbon remineralized
• Inefficient transport of carbon
below thermocline
35 days
7 tonnes FeSO4
150 km2
• Iron addition impacts on
phytoplankton community
structure
• Carbon sequestration efficiency
and remineralization rates
• Shift away from
picophytoplankton
• Majority of carbon fixed was not
remineralized
• Unpublished “massive carbon
export” paper by Smetacek et al.
• Interaction between iron and
phosphorus controls on biological activity in the subtropical
North Atlantic
• Increased N-fixation activity was
observed
• No increase in primary
productivity
• Carbon export not measured
SOFeX-N
SOFeX-S
Coale et al., 2004;
Buesseler et al.,
2004
SERIES
Boyd et al., 2004
EIFEX
Hoffmann et al.,
2006;
Jacquet et al.,
2008
2004
Southern OceanAtlantic
FeeP
Rees et al., 2007;
Karl and Letelier,
2008
2004
Sub-tropical
Northeast
Atlantic-LNLC
21 days
5 tonnes FeSO4
(+20 t PO4)
25 km2
SAGE
Law et al., 2006
2004
Southern Ocean250 km from New
Zealand
15 days
5.4 tonnes
FeSO4
100 km2
• Iron addition’s influence on
sea-air gas exchange
• CO2 drawdown and dimethylsulfide (DMS) production
• Doubling of chlorophyll a but no
significant DMS production and
no significant CO2 drawdown
(preliminary results)
SEEDS-II
Tsuda et al., 2007
2004
Subarctic PacificNorthwest
26 days
491 kg Fe
64 km2
• Monitor ultimate fate of bloom
and carbon for longer time
period than SEEDS-I
• No diatom bloom response
• Increased zooplankton grazing
LOHAFEX
NIO Press Release,
2009
2009
Southern OceanAtlantic
40 days
10 tonnes
FeSO4
300 km2
• Ecological shifts and fate of
sinking carbon
• Increased zooplankton grazing
• Negligible carbon export (preliminary results)
Oceanography
September 2009
243
variable, and small, rates of carbon export
from these iron-induced blooms. But
experiments at this scale cannot resolve
key questions about geoengineering
scenarios for carbon sequestration on
the order of decades to centuries, and on
the scale of the entire Southern Ocean
and beyond. These questions can be
addressed effectively only by using global
biogeochemical models.
From the very beginning of research
on the iron hypothesis, models were used
explicitly to predict what small-scale
experiments could not show: the global
carbon cycle response to large-scale
OIF. In response to the initial interest
in iron fertilization as a potential tool
for global climate mitigation in 1990,
Sarmiento and Orr (1991) modeled
what complete depletion of surface-layer
macronutrients in the Southern Ocean
(by iron-induced phytoplankton blooms)
would do to global atmospheric CO2 and
ocean chemistry. Their model, operating
on a 100-year time frame, predicted
a net global drawdown of around
1–1.5 Gt C yr-1 (98–181 Gt C total over
100 yr). At the time, this figure represented offsetting about 20% of anthropogenic emissions over the 100 years,
corresponding to a delay in the rising
atmospheric CO2 trajectory of about
18 years. Even so, the projections resulted
in a global atmospheric concentration of
CO2 near 700 ppm by 2100. Their model
also predicted a huge area of anoxia in
the southwestern Indian Ocean, and a
high potential for methane production
as a result of this anoxia (Sarmiento and
Orr, 1991; see also Fuhrman and Capone,
1991). Of essential importance is that
these results, by design, represent the
extreme (unrealistic) case scenario—
fertilizing the entire Southern Ocean
244
Oceanography
Vol.22, No.3
with iron for 100 years, and assuming
that all of the macronutrients available
were completely used up. Thus, these
results represent an unachievable (both
logistically and ecologically) upper limit.
Since these initial modeling efforts,
there have been many published
variations on the theme (Table 2)—all
concluding, more or less, that largescale ocean fertilization could, at the
limit, sequester only modest amounts of
carbon relative to global human emissions. All such scenarios involve spatial
scales of the entire Southern Ocean or
concentrated in 16% of the world ocean
would have “dramatic effects on oceanic
ecology which are difficult to predict.”
Fuhrman and Capone (1991) highlighted expected influences of enhanced
nutrient cycling on global production
of the potent greenhouse gases methane
and nitrous oxide. Including tentative assessments of effects on fisheries,
Gnanadesikan et al. (2003) modeled
what some of the long-term downstream
ecological effects of nutrient depletion
might look like, and argued that downstream reduction in productivity could
beyond and time scales of decades to
centuries. Most recently, Zahariev et al.
(2008) modeled how the “elimination
of iron limitation” in the ocean globally
would affect atmospheric CO2. Their
results were similar to previous estimates: rates of 0.9 Gt C yr-1 reduction
in atmospheric carbon, or about 11% of
2004 global emissions, and these sequestration rates could only occur “for a year
or two, even under continuous fertilization.” They conclude that their idealized
model of iron fertilization offers only
a “minor impact on atmospheric CO2
growth” (Zahariev et al., 2008).
The early publications also recognized
the downstream effects of OIF, that
is, (1) the influence of iron-induced
nutrient depletion on surface waters
that are subducted and transported,
and that ultimately resurface elsewhere
with diminished nutrient supplies, and
(2) enhanced decomposition of organic
matter in subsurface waters, depleting
oxygen with ecological and biogeochemical consequences that over time will
extend well beyond the fertilized location. As Sarmiento and Orr (1991) put
it, the fact that an increase in productivity of such a magnitude would be
far outweigh the benefit of the initial
iron-induced carbon sequestration in
terms of a global carbon budget.
In short, marine ecosystems are
complex, and any estimate of the net
impact of iron fertilization on global
carbon storage and greenhouse warming
requires an analysis of all of the downstream potentially negative effects,
including a long-term reduction in ocean
productivity, alteration of the structure
of marine food webs, and a more rapid
increase in ocean acidity (Denman,
2008). As argued by Cullen and Boyd
(2008), such long-term and downstream
effects must not only be acceptably
predictable, but also verifiable, if OIF is
to be considered a viable technology for
climate mitigation. So far, no proponents
of OIF have demonstrated, or even
argued, that these effects can be monitored over decades to reveal statistically
significant assessments against the backdrop of climate variability.
Integrating the Science
and Moving Forward
Between 2005 and 2009, the scientific community reviewed the results
from previous artificial fertilization
Table 2. Selected models of large-scale ocean iron fertilization and its impact on atmospheric carbon inventory.
Most models show high rates of CO2 drawdown in surface waters, with only a small fraction exported below the
pycnocline, and only a small fraction of that exported carbon sequestered. Models converge at < 1Gt C yr-1, and
generally assume near-constant fertilization of vast HNLC areas for decades to centuries.
Overall estimate
of C sequestered
Maximal estimate of
C sequestration rate
Summary
Complete macronutrient depletion due
to iron fertilization of
HNLC regions
98–181 Gt C over 100 yrs
Rates around
1–1.5 Gt yr-1
integrated over
a century
Assumes complete macronutrient depletion due to
OIF of the entire Southern Ocean and results in a 20%
reduction in anthropogenic emissions only if this level
of fertilization is maintained for 100 years.
Gnanadesikan
et al., 2003
Patchy fertilization;
includes downstream
effects of macronutrient depletion on
biological pump
Ultimately, the negative
effect on productivity
from OIF could be
30x the amount of C
exported from OIF
2–20% sequestration of 2 Gt yr-1 as
an initial estimate
of global export
production
Sequestration (for 100 yr) is a small percentage
of annual export production. Overall downstream impacts of OIF may outweigh the carbon
sequestration response.
Aumont and
Bopp, 2006
Uses models based on
OIF experiments to
simulate productivity,
export production, and
ultimately sequestration
70 Gt C over 100 yrs
Export production:
Initial increase
3.8 Gt yr-1, slows
to 1.8 Gt yr-1
Sequestration:
0.3–1 Gt yr-1
Ultimately, 90% of sequestration comes from the
Southern Ocean. Model predicts substantial increases in
productivity. Only a fraction of this productivity is ultimately exported, and only a fraction of that is ultimately
sequestered. Requires constant summer fertilization.
Jin et al., 2008
Models patch to basinscale fertilization for
one decade; analyzes
CO2 drawdown
3.4 Gt C over 10 yrs
N/A
The model shows high atmospheric CO2 uptake efficiency, but low total biological pump efficiency: full
fertilization of the entire Pacific HNLC for 10 yrs results
in 3.4 Gt of CO2 drawdown.
Zahariev et al.,
2008
Complete relief of iron
limitation in the global
ocean
77 Gt C over 5,300-yr
maximum
1 Gt yr-1 maximum
Continuous fertilization of the entire Southern Ocean
results in about 11% offset of global emissions under the
most ideal conditions.
Model
Approach
Sarmiento and
Orr, 1991
experiments and models (de Baar
et al., 2005; Boyd et al., 2007; Powell,
2007–2008) and began a more public
discussion about where the field was, and
should be, going (Boyd, 2008; Buesseler
et al., 2008; Lampitt et al., 2008; Smetacek
and Naqvi, 2008). Although mesoscale
experiments had shown that iron addition caused phytoplankton blooms in
HNLC regions, what happened to the
carbon in those blooms, however, was
less consistent and did not support key
calculations in Martin’s “paleo”-iron
hypothesis. For example, an analysis
of results from experiments in the
Southern Ocean (de Baar et al., 2005)
revealed that the average net dissolved
inorganic carbon (DIC) drawdown
from the atmosphere as a result of
iron enrichment was approximately
4347 mol C per mol Fe. Assuming a
carbon export rate to the deep ocean
of 20% of primary production, carbon
export efficiency (amount of carbon
exported to the deep ocean per unit of
iron added) is thus 870 mol C per mol Fe.
This figure is 200 times less than required
to explain paleo-climate observations
(de Baar et al., 2005). Thus, the collective
carbon export efficiency results from
open-ocean experiments were significantly lower than initially hypothesized.
An important symposium was held
at the Woods Hole Oceanographic
Institution (WHOI) in September
2007 (http://www.whoi.edu/page.
do?pid=14617) that included invited
representatives of private companies
promoting ocean fertilization for climate
mitigation. Experimental results and
modeling predictions, as well as the
policy implications and economics of
OIF, were discussed. In an outgrowth
of that workshop, Powell (2007–2008)
summarized the state of OIF science
as it relates to climate mitigation. He
concluded that iron fertilization may
work in principle to sequester some
carbon, but yields are low and longterm sequestration is difficult to verify,
making it less attractive and making
the cost per ton of carbon sequestered greater than proponents might
hope. Uncertainties about ecosystem
Oceanography
September 2009
245
disruption and other potential downstream negative side effects were also
recognized as a cause for concern.
In an outgrowth of the WHOI
workshop and in an effort to move the
field forward, Buesseler et al. (2008)
focused on the need to reduce uncertainties about OIF as a climate mitigation
strategy, arguing that there is “as yet, no
scientific evidence for issuing carbon
credits from OIF.” They urged a move to
larger and longer experiments because
“ecological impacts and CO2 mitigation
are scale-dependent” (Buesseler et al.,
2008). The proposed experiments would
seek to test a broad array of ecosystem
impacts of iron fertilization, would integrate with modeling efforts to analyze
potential downstream effects of OIF, and
would attempt more detailed measurements of the fate of fixed carbon from
iron-induced blooms.
Around the same time, Smetacek and
Naqvi (2008)—arguing that an apparent
consensus against OIF was premature—
also called for larger and longer experiments to test both the geoengineering
potential of OIF and the potential
for unintended negative side effects.
Collectively, these proposals advocated
cautiously moving forward, gradually
increasing the size of experiments, and
measuring carbon export, net greenhouse gas budgets, food-web disruption,
and other parameters, in order to gain
more information about the potential of
iron fertilization as a climate mitigation
strategy. Following up on arguments
presented at the WHOI workshop,
Cullen and Boyd (2008) pointed out that
it will be necessary to predict and verify
cumulative and long-term effects of OIF
for climate mitigation, and there are
good reasons to believe that it may not
246
Oceanography
Vol.22, No.3
be possible to detect significant downstream effects of wide-scale OIF, such as
increased anoxia and nitrous oxide emissions, against a background of ocean
variability. It follows that if negative
effects were large but masked by natural
variability, they might not be detectable
until they were irreversible. But regardless, it was the responsibility of proponents to show that these effects could be
assessed (Cullen and Boyd, 2008).
As the cycle of experimentation,
media coverage, and calls for more
research on unintended side effects
continued, the scientific questions being
addressed by OIF experiments evolved
substantially. One defining aspect has
emerged: the experiments that were once
focused on controls of ocean productivity and their relationships to climate
are now almost exclusively couched—
either explicitly or implicitly—in terms
of testing iron fertilization as a carbon
sequestration strategy for mitigating
excess global atmospheric CO2.
The first experiment after the 2008
calls for geoengineering research,
LOHAFEX, was conducted in early
2009 amidst international controversy
and after a temporary suspension by the
German government. The controversy
stemmed from a belief that the experiment, the largest and longest conducted
to date, would pave the way for geoengineering projects and, by being an experiment designed to test the idea that iron
fertilization could alter CO2 concentrations, was itself geoengineering.
The initial results of the LOHAFEX
experiment, conducted over 40 days
using 10 tonnes of iron sulfate distributed over 300 km2, demonstrated the
expected iron-induced bloom (AWI
Press Release, 2009b). However, the
bloom also induced an increase in
copepod grazing and amphipod abundance, channeling carbon into the food
web and largely to respiration; in this
way, the vast majority of the newly
fixed carbon was rapidly remineralized and only a negligible amount was
exported. The National Institute of
Oceanography (NIO) report indicated
that these results “dampened the hopes”
of using OIF in the Southern Ocean to
mitigate global climate change (NIO
Press Release, 2009). The controversy
surrounding LOHAFEX brings into
focus the dynamics between OIF science,
companies hoping to conduct OIF
commercially, and international and
national regulations on OIF. No publicly
funded scientific experiments beyond
LOHAFEX have been announced, but
future experiments will be conducted
under new regulatory regimes and,
certainly, under close scrutiny.
O CE AN FERTILIZATION AS A
COMMERCIAL VENTURE
The History of Commercial
Interest Parallels the Science
As the iron hypothesis—and John
Martin’s famous quip—became popularized, so did the idea that fertilizing the
ocean could be a cheap, fast, and easy
solution to the greenhouse gas problem.
Not long after the hypothesis was first
put forward, The Washington Post
published an article on iron fertilization as way to “battle the greenhouse
effect” (Booth, 1990). This attention
would quickly translate into commercial
ventures seeking to profit from ocean
fertilization’s geoengineering potential.
After the first round of press coverage,
the American Society of Limnology and
Oceanography (ASLO) held a workshop
to review the science (ASLO, 1991).
After many presentations and long
discussion, participants agreed on a resolution, endorsed by the Society, “urging
all governments to regard the role of
iron in marine productivity as an area
for further research and not to consider
iron fertilization as a policy option that
significantly changes the need to reduce
emissions of carbon dioxide” (see preface
to Chisholm and Morel, 1991). This
resolution’s emphasis on the scientific
uncertainties surrounding the ecosystem
response to iron fertilization stimulated
important OIF research as intended,
but the carefully worded phrase about
OIF as a policy option did not stem the
growing interest in ocean fertilization for
commercial gain.
Interested in what appeared to be
an easy and intriguing solution for
global warming, the press followed the
IronEx I and II experiments in 1993
and 1995, and interpreted their results
in a different context from the intent of
the experiments. The addition of iron
to equatorial Pacific waters in IronEx I
resulted in a phytoplankton bloom, and
the scientists involved were justifiably
very excited about the results: iron did
indeed stimulate the growth of phytoplankton consistent with the first arm
of the iron hypothesis—that iron limits
productivity in these HNLC waters. The
magnitude of the bloom, however, was
muted, in part because of the design,
and in part because of unexpected
physics during the experiment. The press
coverage of the experiments, however,
focused on the small magnitude of the
bloom and its meaning in the climate
mitigation context, rather than the
exciting fact that there was a bloom
at all. (This view is understandable, as
they are writing for an audience that is
not interested in whether or not iron
limits HNLC regions of the ocean; it is
interested in climate change.) One headline prompted by the low-magnitude
response during the IronEx I experiment
read: “Pumping iron: too weak to slow
warming” (Monastersky, 1994), and
another account reported that “the idea
of fertilizing the entire Southern Ocean
production in the surface ocean.” In this
patent, Markels cited the results of the
IronEx I experiment and suggested that
he could constantly fertilize 140,000 km2
of the Gulf Stream (which is not an
HNLC region) with enough iron, phosphate, and micronutrients to remove
1.3 Gt of CO2 and produce 50 Mt of
additional seafood production annually
(Markels, 1995).
“
We suggest that it is time to break this
two-decade cycle, and argue that we know
enough about ocean fertilization to say that
it should not be considered further as a means
to mitigate climate change.
should probably be considered dead”
(Kunzig, 1994). But such skepticism was
short-lived. When results from IronEx II
were published in 1996, The Washington
Post’s coverage described them as a
confirmation of Martin’s hypothesis of
CO2 sequestration and climatic cooling
(Suplee, 1996).
The First Commercial
Proposal: Ocean Fertilization
for Fish Production
In 1994, before IronEx II had taken
place, Michael Markels Jr., former CEO
of the engineering firm Versar Inc., filed
the first of several patent applications
for a method of improved production of
seafood by ocean fertilization (Markels,
1995). The method referred to fertilizing
with “all nutrients that are found to limit
”
Markels would go on to found Ocean
Farming Inc., a company seeking to capitalize on what it said was ocean fertilization’s promise to increase fish biomass
production. In early 1998, Ocean
Farming Inc. carried out two successive
small-scale (9 km2) ocean iron fertilization “demonstrations” in the Gulf of
Mexico. The results were not published,
but it was reported that, while the iron
fertilization induced an initial increase
in phytoplankton production, the bloom
failed to expand as much as anticipated,
owing to a “second limitation,” most
likely from phosphorus (Markels and
Barber, 2001)—an unsurprising result
because the Gulf of Mexico is a low
macronutrient region, not favorable
to iron-induced bloom development
(Markels and Barber, 2001).
Oceanography
September 2009
247
Citing its own unpublished activities
from the Gulf of Mexico, and the initial
results from IronEx I and II experiments,
Ocean Farming Inc. then reportedly
secured a lease of 800,000 square miles
of LNLC tropical ocean in the exclusive
economic zone of the Republic of the
Marshall Islands in the western Pacific
“
About half of the carbon removed in
fertilized areas will sink to the bottom of
the deep ocean…The continuous fertilization of the same 100,000 square miles
of tropical ocean should sequester about
30 percent of the CO2 produced by the
United States from the burning of fossil
fuels” (Markels, 1998). This statement
But, ocean fertilization research should
not be halted: if used appropriately and
applied to testable hypotheses, it is a powerful
research tool for understanding the
responses of ocean ecosystems in the context
of climate change.
”
Ocean (Markels, 1998). According to the
structure of the lease, once fertilization
of this vast area occurred and fish catch
had commenced, Ocean Farming Inc.
would pay the Marshallese government
$3.75 a square mile or 7% of the profit,
whichever was more (Markels, 1998).
The Change to Carbon Credits
While the fertilization of the waters
near the Marshall Islands never did take
place, the idea of commercializing the
ecosystem response to ocean fertilization
did not disappear. Instead, it morphed.
As Markels wrote in 1998, “Ocean
fertilization also promises benefits that
should be welcomed by those concerned
about possible global warming. The
growth of phytoplankton in the ocean
removes CO2, a greenhouse gas, from
the ocean surface and the atmosphere.
248
Oceanography
Vol.22, No.3
was published before any OIF experiment had provided an estimate of carbon
export from an iron-induced bloom. A
year later in an article in the journal of
the Cosmos Club, Markels discussed
combining the dual purposes of ocean
farming and carbon sequestration from
OIF. He also started a new company,
GreenSea Venture, which sought to
commercialize ocean fertilization to
sell carbon offset credits for sequestered CO2 (Markels, 1999).
A More Crowded Field: GreenSea
Venture Gets a Competitor
In the late 1990s, a second corporation
seeking to invest in ocean iron fertilization as a way to sequester carbon and sell
carbon offset credits, Carboncorp USA,
was created. According to an archived
copy of the Carboncorp USA Web site
(Carboncorp USA, 1999), “Ocean
Carbon Sequestration (OCSTM) patented
nutrient supplements will stimulate an
immediate plankton bloom. This bloom
of plant biomass removes CO2 from the
atmosphere and stores it safely in rich
phytoplankton.” At the time, Carboncorp
USA proposed to use commercial ships
traversing shipping lanes on the high
seas to meter small amounts of the
company’s nutrient supplements into the
water. The idea was to offset the emissions of the shipping by sequestering
carbon and selling credits (Carboncorp
USA, 1999; Adhiya, 2001).
By 2001, Carboncorp USA had
disappeared, and many of its ideas were
presented by Ocean Carbon Sciences
Inc., led by Vancouver-based entrepreneur Robert Falls. Representatives from
both Ocean Carbon Sciences Inc. and
GreenSea Venture were invited participants in a 2001 ASLO workshop on iron
fertilization (ASLO Workshop Statement,
2001), as commercial ocean fertilization
for carbon sequestration was attracting
attention and investment (Krivit, 2007).
Scaling Up: A Proposal For a
Technology Demonstration
Ocean Carbon Sciences Inc. initially
committed $325,000 CAD to the
Canadian component of the SERIES
experiment, but defaulted on the pledge,
did not participate in the experiment
(Canadian SOLAS Report, 2003), and
made little headway toward conducting
its own experiment. GreenSea Venture,
however, was moving forward. In 2001,
Michael Markels teamed up with Richard
Barber, a highly respected biological
oceanographer at Duke University
and one of the scientific team from the
original IronEx experiments. At the 2001
National Energy Technology Laboratory
Conference on Carbon Sequestration,
hosted by the Department of Energy,
Markels and Barber announced plans
for GreenSea to move ahead with a
5,000-square-mile “technology demonstration” of ocean iron fertilization for
carbon sequestration in the HNLC
region of the equatorial Pacific (Markels
and Barber, 2001). Supporting their
proposal, they cited the two IronEx
experiments, SOIREE, and Ocean
Farming’s two Gulf of Mexico demonstrations. Although none of these experiments showed any carbon export from
an iron-induced bloom, it was argued
that increasing the size of the fertilized
patch approximately 100 times from
previous experiments would yield better
estimates of carbon export potential of
OIF (Markels and Barber, 2001). While
the proposed technology demonstration
never did take place, the commercial
potential suggested in their proposal—
quoted at a cost of $2 per ton of sequestered carbon dioxide, including the cost
of verification, overhead, and profit—
served to broaden the discussion about
the ways in which ocean fertilization
experimentation should move forward.
Despite no direct scientific evidence
that ocean fertilization could substantially slow global warming, the simply
stated appeal of OIF as a quick climate
fix continued to attract media attention.
This attention, in turn, further fueled
commercial interest in the idea. In an
interview with Wired magazine in 2000,
Markels claimed that “given 200 boats,
8.1 million tons of iron, and, say, 11
percent of the world’s ocean,” he could
“zero out global warming. Next question?” (Graeber, 2000). The vision of
commercial ventures interested in using
iron fertilization for geoengineering was
unquestionably large in scope and was
not being dictated by the results of initial
scientific OIF experiments.
THE RECENT HISTORY OF
COMMERCIAL OIF AND OIF
REGUL ATION
Commercial Ventures
Continue to Experiment
Scientists Give Mixed Signals
As scientists were gearing up for the
SOFeX experiment in the Southern
Ocean, California-based entrepreneur
Russ George founded the Planktos
Foundation, which billed itself as a
not-for-profit organization seeking to
use iron fertilization to solve global
warming. Planktos offered, for sale by
charitable donation on the Internet,
$4 “Green Tags” (Planktos Green Tags,
2002), which were in effect like personal,
small carbon footprint offsets and would
be used to support the foundation’s
efforts to develop iron fertilization as
a technology for carbon sequestration
(Plotkin, 2002).
In the summer of 2002, temporally
and geographically between SOFeX in
the Southern Ocean and SERIES in the
Gulf of Alaska, the Planktos Foundation
conducted its own iron fertilization
“demonstration,” dumping ironcontaining paint pigment into the North
Central Pacific along a 50-km transect
east of Hawaii (not HNLC waters)
from Neil Young’s antique sailing yacht,
Ragland (Schiermeier, 2003). Journalist
Wendy Williams quoted Russ George
in a piece for Living on Earth as arguing
that this fertilization project had induced
a large phytoplankton bloom and had
sequestered enough carbon to offset the
entire carbon footprint of his California
hometown of Half Moon Bay. George
further described Planktos’ activity as
“really more of a business experiment”
(Williams, 2003). Although the results
of the experiment were never made
public, a description of the activity was
At this point, the activities of GreenSea
Venture and Ocean Carbon Sciences
elicited a response from oceanographers,
who raised concerns about the disconnect between scientific evidence and
commercial proposals. Several of us cited
the difficulties of verifying the magnitude of carbon sequestration explicitly
caused by ocean iron fertilization, and
argued that this would make ocean
fertilization ineligible for carbon credits
(Chisholm et al., 2001). We pointed
out further that widespread ecosystem
disruption is inherent to the design of
ocean fertilization, and the unintended
consequences of this disruption would
likely be significant and unpredictable.
Although we maintained that ocean iron
fertilization should not be conducted
on a commercial scale, we also made
explicitly clear that we were not arguing
against small-scale scientific iron enrichment experiments designed to answer
specific questions about how marine
ecosystems function. Some felt that we
had “overstated the current knowledge
in reaching [our] opinion that iron
fertilization is not a viable option for
CO2 management” (Johnson and Karl,
2002), and countered that verifying
carbon credits is not the critical issue,
what is key is whether OIF “is a feasible
strategy to mitigate increasing CO2 in
the atmosphere” (Johnson and Karl,
2002). Thus, the potential for climate
mitigation using OIF was gaining credibility (at least as a testable hypothesis) in
the scientific community.
Oceanography
September 2009
249
Box 2A. United Nations Convention on the Law of the Sea (UNCLOS)
What:
• International treaty first passed in 1982 that governs territorial legality of the ocean areas.
• United States has not yet ratified it, but many provisions are regarded as binding international law.
• The Law of the Sea sets limits for exclusive economic zones and laws of conduct on the high seas.
• UN General Assembly passes relevant resolutions. UN’s International Maritime Organization and other organizations administer UNCLOS.
How It Relates to OIF: Several general provisions of UNCLOS are relevant to OIF. Article 145 stipulates that use of the high seas for marine
scientific research must be for peaceful purposes geared toward the increase of knowledge and understanding for all humankind. Various
other articles outline requirements of states to protect life of the marine environment. UNCLOS is also relevant to OIF for fish production, as it
governs both the delimitation of fishing rights within exclusive economic zones (EEZs) and fishing rights on the high seas.
Key Statements and Decisions on OIF:
Dec 2007: General Assembly resolution calls for more research into the effects of ocean iron fertilization.
Dec 2008:General Assembly resolution welcomes London Convention and Convention on Biological Diversity
decisions against large-scale OIF.
Future Regulations: Although the General Assembly has not yet passed specific regulatory resolutions regarding OIF, the centrality of
UNCLOS to maritime regulation, and its relevancy to environmental protection, scientific research regulation, and international waters legal
issues, could make it a good forum for coordinated regulation of iron fertilization in the future.
published in the news feature section of
Nature the following year (Schiermeier,
2003), including a photo of George on
Ragland. The Planktos Foundation, and
the entire idea of carbon credit sale from
ocean fertilization, was gaining attention
and momentum.
Legal Issues Arise
As it became evident that commercial
operations were moving forward,
concerned scientists and environmental
groups continued to raise questions
about the efficacy and legality of
commercial ocean fertilization. There
appeared to be no law preventing
ocean fertilization beyond the 200-mile
exclusive economic zone of any country
(Markels and Barber, 2001; McKie,
2003). Indeed, to our knowledge, the
250
Oceanography
Vol.22, No.3
initial, small-scale, scientific OIF experiments were carried out without permits,
as there were no clear statutes at the
time. Also, it was widely recognized
that the small scale of the experiments
rendered their impacts miniscule and
ephemeral by any measure.
Several international treaties cover
activities in international waters,
including the United Nations Convention
on the Law of the Sea, the 1972 London
Convention on Marine Pollution and
Ocean Dumping, and the 1959 Antarctic
Treaty, which covers portions of the
Southern Ocean. The UN Convention
on Biological Diversity also regulates
activities that will impact ecosystem
species diversity (see Boxes 2A, 2B, and
2C). At a 2001 ASLO workshop on ocean
fertilization, the former secretary of
the Intergovernmental Oceanographic
Commission, Geoff Holland, stated
that “there are no legal precedents that
apply directly to ocean fertilization,
though there are parts of existing laws
that are relevant” (Holland, 2001). At
the time, it was not clear whether ocean
fertilization fell under the purview of
the London Convention, and the lack
of clarity on the legal issue meant that
commercial interests could move ahead,
for the time being.
Commercial Ventures Advance
Plans for Fertilization as Scientists
Debate Its Future
After 2004, while scientists were
analyzing previous results and planning
future experiments, commercial interests
continued to advance their plans. The
Box 2B. Convention on the Prevention of Marine Pollution by Dumping of Wastes
and Other Matter (1972 London Convention and 1996 London Protocol)
What:
• 1972 intergovernmental treaty involving 86 countries.
• Headquarters: International Maritime Organization in London, a UN organization.
• London Convention regulates pollution via dumping at sea (but not coastal runoff).
• Some dumping pollution (such as scrap metal) is regulated, and some (such as mercury) is banned outright.
• 1996 London Protocol expands the London Convention to use the “Precautionary Principle” and reduce pollution.
• United States is party to Convention, but has signed but not ratified the London Protocol.
• US enforcement of London Convention by EPA under Marine Protection, Research, and Sanctuaries Act of 1972.
• Enforcement is by each country in its own waters and to ships sailing under its flag. No international enforcement.
How It Relates to OIF: Because iron fertilization activities require the addition of inorganic material in the HNLC regions of the world
ocean, and generally take place in international waters, the London Convention is deemed relevant. This relevancy has been the subject of
some discussion, however, as the London Convention regulates dumping for disposal of waste. Because OIF for carbon sequestration would
sequester carbon in the deep ocean, the discussion of whether OIF is more appropriately the dumping of carbon has also been raised.
Key Statements and Decisions on OIF:
Jun 2007: Scientific Group issue “Statement of Concern” to the London Convention over large-scale ocean fertilization activities.
Nov 2007: London Convention Resolution adopts “Statement of Concern” and agrees that ocean fertilization falls under the resolution’s
purview.
May 2008: Scientific Group outline what ocean iron fertilization entails and detail some unknown risks.
Oct 2008: London Convention Resolution states that no ocean fertilization activities other than “legitimate scientific research” (to be defined
in 2009) should be allowed.
Feb 2009:Technical Working Group drafts ecological risk assessment framework for OIF research stipulating that the research must be
hypothesis-driven. This point does not rule out commercial gain or carbon credits from legitimate research.
Feb 2009: Legal Working Group proposes new resolutions and new binding articles on OIF, outlining future regulatory possibilities for
consideration of the full Convention.
May 2009: Scientific Group meet in Rome to consider reports from the working groups and to start outlining the acceptable levels of negative environmental impacts from legitimate scientific research (i.e., experiments 200 x 200 km or smaller).
Future Regulation: The London Convention is clearly grappling with the difficult issue of regulating scientific research at the supranational level. At the same time, there is a clear push toward further restriction of nonscience OIF activities. Specific proposals from the February
2009 working groups range from outright bans of all nonscience OIF activities to “suspensions” of activity until more data can be gathered.
The questions of commercialization, ecological assessment, and degree of enforcement/intensity of regulation have all been raised but none
of them is resolved. The Scientific Group have started to discuss the questions about how much of an environmental impact from scientific
research is too much and will present a report to the governing bodies of the convention when they meet in London in the fall of 2009.
Planktos Foundation became Planktos
Inc. in 2005 and was shortly thereafter
purchased by Solar Energy Limited
(Business Wire, 2005a). Later that year,
Diatom Corporation agreed to purchase
the marketing rights for carbon credits
generated from Planktos Inc.’s iron
fertilization activities (Business Wire,
2005b). About 18 months later, publicly
traded Diatom Corp. became Planktos
Corp. (Business Wire, 2007). During
this time, the other commercial entity
involved in OIF, GreenSea Venture,
had moved from ocean fertilization
demonstrations to financial support for
Oceanography
September 2009
251
Box 2C. United Nations Convention on Biological Diversity (CBD)
What:
• International treaty established at the Earth Summit in Rio de Janeiro in 1992.
• Dual purposes of biodiversity conservation and sustainable use and equitable distribution of resources.
• All UN nations are party, except the United States, which has signed but not ratified the treaty.
• The Convention is administered by the UN Environment Programme (UNEP).
• Enforcement by individual ministries of member states.
How It Relates to OIF: The Convention regulates actions that threaten biodiversity, including marine biodiversity. In this way, the
unknown ecological effects of large-scale OIF implementation fall under the purview of the Convention.
Key Statements and Decisions on OIF:
May 2008 CBD Decision: “All large-scale OIF activities should not be allowed.”
• Makes exception for “small scale studies in coastal waters.” Coastal waters may be an aberration.
• Decision explicitly mentions commercial interests as a reason not to allow OIF.
• Decision passed because of ecological risks of OIF and uproar over Planktos’ experimental plans.
• Widely viewed as a “UN moratorium” on commercial OIF activities.
Future Regulations: The CBD suggested that OIF regulation be done in coordination with the International Maritime Organisation
(IMO) and London Convention. It calls for a global transparent control and regulatory mechanism to be established, and urges consultation
with all parties involved to establish a knowledge base about the associated ecological risks.
Note on CBD Enforcement: In early 2009, the German research ministry cited the UN CBD COP 9 moratorium on ocean fertilization
in its decision to suspend the Indo-German LOHAFEX OIF experiment. It demanded more environmental risk assessments and independent
scientific assessments of the project, specifically mentioning the coastal water stipulation and citing the ecological concerns raised by CBD.
Although the experiment was eventually given the green light, it was a first test of country-specific enforcement of international treaties on this
issue and has informed the London Convention working groups’ discussions on how best to regulate the science of OIF.
biogeochemical modeling (Chu et al.,
2003; Rice, 2003). Meanwhile, Dan
Whaley, a former information technology entrepreneur, founded Climos
Inc., a for-profit company setting out to
pursue the same goals as Planktos Inc.:
selling carbon credits from ocean iron
fertilization (Riddell, 2008). Thus, by
2006, two companies operating in the
San Francisco Bay area—Climos and
Planktos—were developing the business
of iron fertilization.
In late 2006, it was reported that
Planktos donated carbon credits to two
252
Oceanography
Vol.22, No.3
California environmental organizations
to make them “carbon neutral” for 2007
(Industrial Environment, 2006), a claim
that was based on an upcoming pilot
demonstration of ocean fertilization.
At this time, Planktos was describing
itself as an “ecorestoration” company,
seeking to replenish what was described
as diminished plankton stocks in the
ocean, in addition to sequestering carbon
for offset credits. In a 2007 presentation
to the US House Committee on Energy
Independence and Global Warming,
Planktos’ Russ George stated that “our
ocean plankton restoration pilot projects
will generate the first substantial iron
seeded blooms aimed at serving our twin
purposes of restoring ocean plant ecosystems and sequestering atmospheric CO2”
(George, 2007).
Planktos Folds Before
It Can Fertilize
In early 2007, Planktos Inc. announced
plans to conduct a large fertilization
experiment in the equatorial Pacific near
the Galápagos Islands. Like the proposed
Markels-Barber demonstration, the
experiment was to be on a previously
unachieved scale of 10,000 km2; it would
be the first pilot project in a planned
“Voyage of Recovery” (see Figure 5
for a sense of the relative size of this
experiment). The company purchased a
retired research vessel, Weatherbird II,
from the Bermuda Biological Station to
execute the experiment (Environmental
Protection Agency, 2007). The scale and
seriousness of this proposal, coupled
with increasing concerns about risks to
marine ecosystems, mobilized environmental NGOs. The Canadian-based ETC
Group called on the US Environmental
Protection Agency (EPA) to stop the
Planktos experiment (ETC Group News
Release, 2007). The World Wildlife
Fund argued that Planktos was taking
too big and too dangerous a leap with
the scale and intent of the experiment,
especially because it was near the
Galápagos Islands (Sullivan, 2007). The
Sea Shepherd Society threatened to block
Planktos’ fertilization ship physically
(South Bay, 2007), arguing that it was a
violation of international laws on marine
dumping (the London Convention; Sea
Shepherd News, 2008).
With international law once again
in question, the legal arguments began
to mount. The Scientific Group of the
London Convention issued a June 2007
statement of concern about Planktos’
plan (Scientific Group of the London
Convention Meeting Report, 2007). In
a statement submitted at that London
Convention meeting, the US EPA
announced that it would not permit
the Planktos plan to proceed under the
US Marine Protection Research and
Sanctuaries Act of 1972 (the US enforcement of the London Convention) if
Planktos was flying a US flag from
Figure 5. The relative sizes of fertilized patches in ocean fertilization experiments, and in demonstrations
carried out, or proposed, by the private sector. At bottom is an estimate (based on the model of
Zahariev et al., 2008) of the size of the patch that would result from fertilizing the HNLC regions of
the Southern Ocean.
Weatherbird II (Parks, 2008; Scientific
Group of the London Convention
Meeting Report, 2007). Apparently,
Planktos assured the EPA that it would
not be sailing under a US-flagged ship
(International Center for Technology
Assessment, 2007), which seemed to
contradict that Weatherbird II was a
US-registered vessel and would set out
from a US port for the experiment.
Because of the proposed experiment’s
proximity to the Galápagos, the government of Ecuador also issued statements
of alarm (Scientific Group of the London
Convention Meeting Report, 2007).
In the fall of 2007, as Planktos’ vessel
prepared to leave the eastern coast of
the United States, the full Conference of
Parties to the London Convention issued
a statement of concern about the legality
and wise practice of large-scale ocean
iron fertilization activities, taking the
first step toward explicit international
regulation of iron fertilization (London
Convention Meeting Report, 2007).
Planktos’ experiment did not take
place (Courtland, 2008). Citing concerns
over disruption of their activities,
Weatherbird II left port headed to an
undisclosed location in the Atlantic
Oceanography
September 2009
253
Ocean. When the ship approached the
Spanish-owned Canary Islands, reportedly to pick up supplies, the Spanish
government refused it port entry
(Consumer Eroski, 2007), a decision
“
Since its inception, Climos made clear its
intention to conduct research in collaboration with the scientific community,
and to this end brought in Margaret
Leinen, an accomplished oceanogra-
The potent intellectual resources that
have been tied up in the ocean fertilization
controversy year after year should be freed
to pursue more effective responses to the
pervasive threat of carbon dioxide emissions
and global warming.
”
that Planktos blamed on environmental
organizations (Thompson, 2008). The
ship eventually docked in the Portuguese
island of Madeira to take on needed
supplies (San Francisco Business Times,
2007). Investors pulled out of Planktos,
and, in early 2008, Planktos Corp. ended
its relationship with Planktos Inc., as
did Solar Energy Limited. Planktos Inc.
then suspended operations (Business
Wire, 2008; Kerry, 2008), citing both
a lack of funds and a “highly effective
disinformation campaign waged by
anti-offset crusaders” (Brown, 2008).
Several months later in June 2008, Russ
George started a new iron fertilization “ecorestoration” company, named
Planktos-Science, though there has been
little activity other than Web posts from
this company thus far (see http://www.
planktos-science.com/).
Climos’ Path Forward
These events left Climos as the leading
company involved in the nascent
commercial iron fertilization business.
254
Oceanography
Vol.22, No.3
pher, and former Assistant Director for
Geosciences at the US National Science
Foundation, to be Chief Science Officer
(Climos About Us, 2008). Climos’ initial
plan was to attract substantial business
investment and to employ environmental consulting firms to produce
codes of conduct for iron fertilization
experiments as a first step toward
eventual verification for carbon credit
sale (Climos Press, 2007). Climos representatives have attended international
meetings, UN meetings, and scientific
conferences to participate in debating
the future of iron fertilization science.
At the same time, the company has also
made clear its plans to conduct its own
technology demonstration (Murray,
2008). In early 2008, Climos issued a
press release responding to Greenpeace
criticisms of plans for commercial OIF
(Allsopp et al., 2007). In particular,
Climos challenged what it described
as Greenpeace’s assumption that OIF
for carbon sequestration would require
large-scale and continuous fertilization,
something that “no commercial entity
has suggested should take place before
a period of experimentation” (Leinen
et al., 2008). On its Web site and in the
press, Climos has announced plans for a
Southern Ocean demonstration experiment that would be “part of a new phase
of research focused on the efficacy and
impact of moderately sized experiments
(< 200 x 200 km)” that would “emphasize
research related to export and sequestration as well as environmental impact”
(Climos FAQ, 2008). In a statement to
the press, Climos officials indicated that
they hope to conduct their first trial by
the end of 2009 and to be able to start
selling carbon credits shortly thereafter
(Murray, 2008). More recently, Leinen
has been identified as Chief Executive
Officer of the Climate Response Fund,
“a nonprofit organization formed to
provide funding and support for other
activities needed to explore innovative solutions to the effects of climate
change” (Climate Response Fund, 2009).
At the time of this writing, it is unclear
whether the formation of this new
nonprofit organization will influence
the plans of Climos.
Regulatory Clarity Is
On The Horizon
In part due to the activities of Planktos,
Climos is operating in a vastly changed
regulatory arena than existed in the
early years of commercial interest in
ocean fertilization. Planktos’ proposed
experiment, and the vigorous response
to it from environmental NGOs and the
London Convention regulators, showed
the first signs of a negative feedback
loop. Up until this point, the trajectory had been continued expansion
of commercial interest and proposed
demonstrations, despite the absence of
direct evidence for commercial claims
from the results of scientific experiments. After concerns were expressed at
the 2007 London Convention assembly
regarding the planned Planktos experiment in fall 2007, representatives met
again in May 2008. They reiterated their
concerns about unchecked large-scale
fertilization activities, and defined
how iron fertilization fit under the
jurisdiction of the London Convention
(Scientific Group of the London
Convention Meeting Report, 2008).
That same month, members of the
UN Convention on Biological Diversity
passed a decision on iron fertilization
(UN CBD, 2008), citing the London
Convention’s statements of concern.
They requested all member states to
ensure that ocean iron fertilization
activities do not take place, with the
exception of small-scale scientific studies
in coastal waters (see below), until there
is adequate scientific basis on which to
justify these activities. They emphasized
that the excepted small-scale studies
could not be used for the generation of
carbon offset credits.
While the Convention on Biological
Diversity did not explicitly define what
was meant by “small scale,” a report
submitted by the Intergovernmental
Oceanographic Commission (IOC), part
of UNESCO, to the London Convention
Scientific Group at Guayaquil in 2008,
proposed that OIF activities greater
than a size of 200 x 200 km (40,000 km2)
should be considered “large scale,” and
anything smaller would remain “small
scale” (Scientific Group of the London
Convention Meeting Report, 2008).
The area 200 x 200 km is two orders of
magnitude larger than the largest OIF
experiments to date (Figure 5).
The stipulation of “coastal waters” in
the CBD statement was largely viewed
as an aberration, as most coastal waters
are not iron limited, and this is why
all previous scientific studies have
taken place in the open ocean (Owens,
2009; Alfred Wegener Institute, 2009).
Nonetheless, because the stipulation
was written into the UN CBD language,
members of the LOHAFEX scientific
team found themselves having to find a
way to describe their open-ocean experiment as “coastal” in order to proceed.
multi-step delineated environmental risk
assessment framework was proposed
for future experiments, including smallscale scientific experiments (Report of
the First Meeting of the Intersessional
Technical Working Group on Ocean
Fertilization, 2009). The hope is to avoid
the kind of controversy and uncertainty
surrounding LOHAFEX by having a
system in place to prevent confusion
between scientific experiments and
geoengineering projects.
In May 2009, the Scientific Group of
the London Convention met to discuss
Thus, when providing their risk assessment to the German Research Ministry
during the experiment’s suspension (see
introduction above), they argued that
iron fertilization in the Southern Ocean
would stimulate the growth of “coastal
species” of phytoplankton, an argument
that was apparently accepted, as the
German ministry ultimately allowed the
experiment to proceed. If science and
policy are to move forward on this issue,
lines of communication between sectors
should be continually improved.
A few months after the Convention
on Biological Diversity decision, the full
London Convention took up the issue of
ocean iron fertilization once more, and
passed a Resolution on the Regulation
of Ocean Fertilization, announcing that
all iron fertilization activities, with the
exception of “legitimate science,” were
in violation of the London Convention’s
regulations on marine dumping. The
London Convention also decided that
technical and legal working groups
would meet in February 2009 to
determine what constitutes legitimate
scientific research, establish an assessment framework, and propose future
regulations under the Convention. A
the report of the Technical Working
Group and attempt to further define the
acceptable ecological impacts of “legitimate scientific research.” Specifically on
the agenda was a “Draft Action List for
Ocean Iron Fertilization,” proposed by
representatives from Australia and New
Zealand, which lists upper and lower
limits for nutrient and chemical species
concentrations (dissolved oxygen, pH,
nitrous oxide, ammonium, and methane,
among others) altered in response to
a small-scale (less than 200 x 200-km)
scientific experiment. Draft Action Lists
are used by the London Convention
process to regulate substances on the
basis of their effects on the marine
environment; the 2009 version entitled
“Ocean Fertilization: Development of a
Draft Action List” is the first step in the
process of answering the question, “How
much of a negative result from OIF is too
much?” In the fall of 2009, the governing
body of the London Convention will
take up the reports from the meeting of
the Scientific Group.
What remains unclear is whether
these assessments of future small-scale
scientific OIF experiments will be
conducted at the international level
Oceanography
September 2009
255
of the London Convention, or by the
science and research entities of the
individual countries involved. Future
regulation of larger-scale and nonscientific OIF activities also remains unclear.
The Legal Working Group of the London
Convention proposed a set of possible
regulatory frameworks, ranging from a
simply structured nonbinding resolution against large-scale OIF, to a full
article amendment to the Convention
explicitly prohibiting it (Report of the
First Meeting of the Intersessional Legal
and Related Issues Working Group on
Ocean Fertilization, 2009). The working
groups also avoided resolving the issue of
whether commercialized carbon credits
could be sold as a product of “legitimate
scientific research,” leaving that question
to the full Convention. What is now
clear is that the London Convention is
the central international legal framework for this issue, and that regulation
of OIF will be a subject of continued
discussion and debate.
WHAT HAVE WE LE ARNED?
A Repeated Cycle
Looking back, it is obvious that the
first few years of debate about the iron
hypothesis were a condensed version
of the 20 years to follow. Small-scale
ocean fertilization experiments were
executed by oceanographers to address
specific scientific questions; far-reaching
interpretations, conclusions, and ramifications were highlighted by the media
and by entrepreneurs, prompting a
response from the scientific community
urging caution, followed by calls from
academics and commercial-sector
proponents of OIF for more research.
This cycle has played out repeatedly as
the field of ocean iron fertilization has
256
Oceanography
Vol.22, No.3
moved forward. And, despite the absence
of direct scientific evidence for OIF as
an effective long-term climate mitigation
tool, appeals for research on its efficacy
for this purpose have intensified over
the 20 years since the first experiments,
leading to larger and longer experiments.
The dampening feedback from negative
results, central to the advancement of
scientific research, seems largely absent
from this cycle.
The Science Does Not Support
OIF For Global Geoengineering
Nearly two decades of scientific OIF
experiments have taught us that iron
limits productivity in several regions
of the ocean. We have learned that the
carbon export response to OIF is highly
variable—strongly regulated by the availability of light and silicate as influenced
by physics, and also sensitive to the
interplay of many other factors that have
yet to be resolved. And we have learned
that on the temporal and spatial scales
of the experiments, carbon export is
often quite small due to remineralization of the phytoplankton bloom in the
surface waters. Models show that at the
limit—assuming complete macronutrient depletion and fertilization of the
Southern Ocean for 100 years—what
could be expected at most is global
sequestration of 1.0 Gt C yr-1 (Table 2).
In all modeled scenarios, the amount
of CO2 sequestered is small relative to
the amount predicted to be released by
fossil fuel burning, and the estimates
of sequestration have only gotten
smaller with more experimentation and
modeling (Denman, 2008). Furthermore,
models show that in order to maintain
the carbon sequestration, we would
have to continue to fertilize the entire
Southern Ocean with enough iron to
deplete macronutrients, in perpetuity.
This scenario is not realistic.
Uncertainty and Risk
While the uncertainties about the
efficacy of carbon sequestration from
OIF as a geoengineering proposal
are high (Buesseler et al., 2008), the
certainty of ecological disruption is also
high. OIF for carbon sequestration is
designed to initiate a floristic shift to the
production of larger, bloom-forming
phytoplankton—in particular, diatoms
that are heavy and can sink rapidly. This
fundamental alteration of the base of a
food web would change the structure
and biogeochemical function of the
community that depends upon it. The
induced blooms would consume the
excess macronutrients at the surface
in HNLC regions, which, in combination with enrichment of deep waters
(e.g., Fuhrman and Capone, 1991),
would over time alter the biogeochemistry of the global ocean ecosystem.
Although we cannot predict the precise
changes that would occur, the only way
in which iron fertilization can work for
climate mitigation is to change deep
ocean chemistry and the way endemic
marine food webs function. There is
considerable risk in trusting inherently
uncertain predictions of such large-scale
and long-term alterations of the ocean.
Long-Term Ocean Carbon
Sequestration from Iron
Fertilization Is Not Verifiable
Fertilization-induced changes in
ecosystem function would not only
have profound effects on the ecology of
huge marine ecosystems, but they would
also affect the potential efficacy of OIF
as a carbon sequestration strategy. As
Gnanadesikan et al. (2003) point out,
in order to accurately model the net
global benefit of carbon sequestration,
all downstream effects on biological
productivity must be counted, including
potential disruption of fisheries from
the depletion of macronutrients in the
source waters of productive ecosystems.
These negative effects may outweigh
the benefits of carbon export. Besides
the potential for changes in net global
productivity, OIF could stimulate
nitrous oxide production as a result of
increased remineralization of carbon
and nitrogen (Denman, 2008). This
would result in longer-term and far-field
changes in nitrous oxide production
that could potentially offset significant
amounts of predicted green-house gas
benefits of OIF (Law, 2008). It can thus
be argued that measurements associated
with individual experiments cannot
be adequate to verify what would
ultimately be a long-term, large-scale
effect of many applications of iron
(Cullen and Boyd, 2008).
These complex downstream responses
to ocean fertilization make verification of net greenhouse gas reduction
through fertilization next to impossible
(Chisholm et al., 2001; Cullen and Boyd,
2008; Gnanadesikan and Marinov, 2008).
Furthermore, carbon export measured as
a result of a fertilization-induced bloom
would have to be referenced to a baseline
rate of carbon export. As ocean fertilization and carbon flux research has shown,
this natural rate of carbon export is
highly variable in space and time; establishing an appropriate baseline to grant
carbon credits for individual applications
would be exceedingly difficult, and rife
with uncertainty (Figure 6).
At present, there is no system under
the Kyoto Protocol’s Clean Development
Mechanism to provide for carbon
credits from offsets by marine carbon
sequestration (Powell, 2007–2008).
Thus, under the current international
mechanism, any credits granted
would therefore have to be sold on
the currently unregulated “voluntary
carbon credit market.” Although Climos
Inc. submitted a carbon sequestration
methodology to Det Norske Veritas, an
international verification company, in
late 2007 (Climos Press, 2007), official
approval or verification has not been
given to OIF as a carbon sequestration
methodology. Nonetheless, in the future,
international carbon credit regulatory
systems may well include provisions for
marine “sequestration” offsets (Powell,
2007–2008). It will then be very important for OIF proponents to show in their
plans that the effects of wide-scale, longterm carbon sequestration from OIF
are predictable, acceptable, and statistically verifiable across ocean basins over
decades. In our opinion, assessments of
individual applications are not enough.
Should We Continue to Test OIF
for Climate Mitigation?
Arguments have been made for 20 years
that OIF should not be pursued as a
“quick-fix” for the climate problem. We
hope we have shown that the original
arguments have not been weakened
Figure 6. Commercial ocean fertilization, which would involve validated carbon credits,
would have to demonstrate that the iron-induced bloom would not have otherwise
occurred. This satellite chlorophyll a image shows the bloom induced by iron addition
during LOHAFEX (red circle). Three other blooms are circled in black. How would one
prove that the intentionally fertilized waters, if left alone, would not have bloomed like the
neighboring waters? Image taken from: http://www.awi.de/en/infrastructure/ships/polarstern/
weekly_reports/all_expeditions/ant_xxv/ant_xxv3/26_february_2009/
Oceanography
September 2009
257
by new evidence over that interval.
Ocean fertilization will not solve the
CO2 problem, and if implemented for
profit, regardless of scale, has the potential to change the nature of the ocean
through the “tragedy of the commons”
(Hardin, 1968). Perhaps it is time to
break the two-decade cycle of debate
and accept that we know enough about
ocean fertilization to say that it should
not be considered further as a means
of climate mitigation. But our opinions
aside—and they are indeed opinions,
though science-based—a fundamental
issue remains: the ecological and climate
mitigation response to OIF is scaledependent (Buesseler et al., 2008; Cullen
and Boyd, 2008), and the biogeochemical
changes would be cumulative. The only
way to test OIF as a climate mitigation
tool (i.e., to see if model projections
are right) would be to alter much of
the ocean system, perhaps irreversibly,
before crucially important negative
effects could be evaluated with statistical
confidence. We feel that the risk of doing
this does not compare well to even the
most optimistic predictions of potential
climate mitigation.
Moving On
Climate change is already upon us.
Society needs to know what the ocean
will be like in a high CO2 world—which
ecosystems will be at risk as the ocean
warms and acidifies, and how the altered
ocean will in turn influence climate.
Understanding how ocean biogeochemical cycles are linked, and the processes
that drive these cycles, is essential for
climate prediction. Transformational
developments in genomics along with
rapid advances in ocean observations and
modeling (Doney et al., 2004) allow us
258
Oceanography
Vol.22, No.3
to study the ocean as a system on scales
from molecules to ocean basins, and we
are poised to fully integrate studies of
evolution and biogeochemistry (Woese
and Goldenfeld, 2009). These developments open new vistas for understanding
the very basis of life processes, and it is
these very life processes that, scaled up,
regulate the concentration of CO2 and
other greenhouse gases in Earth’s atmosphere. In this context, small-scale ocean
perturbation experiments are useful
tools for probing ecological and biogeochemical relationships, and they should
continue to be used as such. The potent
intellectual resources that have been tied
up in the ocean fertilization controversy
year after year should be freed to pursue
more effective responses to the pervasive threat of carbon dioxide emissions
and global warming.
Acknowled gments
ALS and SWC are supported in part
by funds from the Gordon and Betty
Moore Foundation Marine Microbiology
Initiative, the US National Science
Foundation, and the US Department
of Energy. JJC is funded by the Natural
Sciences and Engineering Research
Council of Canada, the Killam Trusts,
and the Office of Naval Research. None
of the authors have solicited or received
funds from organizations or companies
for or against ocean fertilization. We
thank Jim Long for graphics, Edward
Boyle for his expert advice on one
part of the manuscript, Charles Miller
for extremely helpful comments and
discussion throughout, and E. Virginia
Armbrust, Heidi Sosik, Kenneth
Denman, and an anonymous reviewer
for their careful and insightful reviews
of the manuscript.
Reference s
Adhiya, J. 2001. Is Ocean Fertilization a Good
Carbon Sequestration Option? Master’s Thesis,
Massachusetts Institute of Technology.
Alfred Wegener Institute. 2009. Risk Assessment
for LOHAFEX. Bremerhaven, Germany, 19 pp.
Available online at: http://www.awi.de/fileadmin/
user_upload/News/Selected_News/2009/
LOHAFEX/0%20AWI_NIO_LOHAFEX_Risk_
Assessment.pdf (accessed August 2, 2009).
Allsopp, M., D. Santillo, and P. Johnston. 2007. A
scientific critique of oceanic iron fertilization as
a climate change mitigation strategy. Available
online at: www.greenpeace.to/publications/iron_
fertilisation_critique.pdf (accessed August 6, 2009).
Anderson, R.F., M.Q. Fleisher, Y. Lao, and G. Winckler.
2008. Modern CaCO3 preservation in equatorial
Pacific sediments in the context of late-Pleistocene
glacial cycles. Marine Chemistry 111:30–46.
Anderson, R.F., G. Winckler, and M. Fleisher. 2007.
What can we learn from the paleo record about
past changes in ocean productivity and controls of
atmospheric CO2? Paper presented at the Woods
Hole Oceanographic Institution Iron Fertilization
symposium, Woods Hole, MA, September 26–27,
2007, http://www.whoi.edu/page.do?pid=14617.
ASLO. 1991. American Society of Limnology and
Oceanography Symposium Report on What Controls
Phytoplankton Production in Nutrient-Rich
Areas of the Open Sea. San Marcos, California,
February 22–24, 1991, 21 pp.
ASLO Workshop Statement. 2001. The Scientific and
Policy Uncertainties Surrounding the Use of Ocean
Fertilization to Transfer Atmospheric Carbon
Dioxide to the Ocean. Washington, DC, April 25,
2001. Available online: http://also.org/docs/ocnfert.
html (accessed August 6, 2009).
Assmy, P., J. Henjes, C. Klaas, and V. Smetacek. 2007.
Mechanisms determining species dominance in a
phytoplankton bloom induced by the iron fertilization experiment EisenEx in the Southern Ocean.
Deep-Sea Research Part I 54(3):340–362.
Aumont, O., and L. Bopp. 2006. Globalizing results
from ocean in situ iron fertilization studies.
Global Biogeochemical Cycles 20, GB2017,
doi:10.1029/2005GB002591.
AWI News. 2009. LOHAFEX. Available online at:
http://www.awi.de/en/news/selected_news/2009/
lohafex/ (accessed August 2, 2009).
AWI Press Release. 2009a. Polarstern expedition
“LOHAFEX” can be conducted. January 26, 2009.
Available online at: http://www.awi.de/en/news/
press_releases (accessed March 25, 2009).
AWI Press Release. 2009b. Lohafex provides new
insights on plankton ecology: Only small amounts
of atmospheric carbon dioxide fixed. March 23,
2009. Available online at: http://www.awi.de/en/
news/press_releases (accessed August 1, 2009).
Booth, W. 1990. Ironing out the “greenhouse effect”:
fertilizing oceans is proposed to spur algae. The
Washington Post. May 20, 1990.
Boyd, P.W. 2008. Implications of large-scale iron
fertilization of the oceans: Introduction and
synthesis. Marine Ecology Progress Series: Theme
Section 364:213–218.
Boyd, P.W., T. Jickells, C.S. Law, S. Blain, E.A. Boyle,
K.O. Buesseler, K.H. Coale, J.J. Cullen,
H.J.W. de Baar, M. Follows, and others. 2007.
Mesoscale iron enrichment experiments
1993–2005: Synthesis and future directions.
Science 315(5812):612–617.
Boyd, P.W., C.S. Law, C.S. Wong, Y. Nojiri, A. Tsuda,
M. Levasseur, S. Takeda, R. Rivkin, P.J. Harrison,
R. Strzepek, and others. 2004. The decline and fate
of an iron-induced subarctic phytoplankton bloom.
Nature 428(6982):549–553.
Boyd, P.W., A.J. Watson, C.S. Law, E.R. Abraham,
T. Trull, R. Murdoch, D.C.E. Bakker, A.R. Bowie,
K.O. Buesseler, H. Chang, and others. 2000. A
mesoscale phytoplankton bloom in the polar
Southern Ocean stimulated by iron fertilization.
Nature 407(6805):695–702.
Brown, S.E.F. 2008. Planktos cancels ocean project,
recalls ship, cuts jobs. San Francisco Business Times.
February 13, 2008. Available online at: http://
www.bizjournals.com/sanjose/stories/2008/02/11/
daily49.html (accessed August 3, 2009).
Buesseler, K.O. 1998. The decoupling of production
and particulate export in the surface ocean. Global
Biogeochemical Cycles 12:297–310.
Buesseler, K.O., L. Ball, J. Andrews, J.K. Cochran,
D.J. Hirschberg, M.P. Bacon, A. Fleer, and
M. Brzezinski. 2001. Upper ocean export of particulate organic carbon and biogenic silica in the
Southern Ocean along 170°W. Deep-Sea Research
Part II 48:4,275–4,297.
Buesseler, K.O., J.E. Andrews, S.M. Pike, and
M.A. Charette. 2004. The effects of iron fertilization on carbon sequestration in the Southern
Ocean. Science 304(5669):414–417.
Buesseler, K.O., S.C. Doney, D.M. Karl, P.W.
Boyd, K. Caldeira, F. Chai, K.H. Coale,
H.J.W. de Baar, P.G. Falkowski, K.S. Johson, and
others. 2008. Environment: Ocean iron fertilization: Moving forward in a sea of uncertainty.
Science 319(5860):162.
Business Wire. 2005a. Solar announces the acquisition
of Planktos, Inc. Business Wire, August 12, 2005.
Available online at: http://www.allbusiness.com/
legal/environmental-law-air-quality-regulation/
5037838-1.html (accessed August 2, 2009).
Business Wire. 2005b. Solar’s subsidiary, Planktos,
enters into funding agreement with Diatom in
exchange for exclusive intellectual property and
marketing rights. Business Wire, September 13,
2005. Available online at: http://www.redorbit.com/
news/science/238992/solars_subsidiary_planktos_
enters_into_funding_agreement_with_diatom_in/
index.html (accessed August 1, 2009).
Business Wire. 2007. Diatom announces name
change, forward stock split and symbol change.
Business Wire, March 8, 2007. Available online
at: http://findarticles.com/p/articles/mi_m0EIN/
is_2007_March_8/ai_n27307673/ (accessed
August 3, 2009).
Business Wire. 2008. Planktos shareholder update.
Business Wire, August 20, 2008. Available online at:
http://www.businesswire.com/portal/site/home/
permalink/?ndmViewId=news_view&newsId=
20080820005832&newsLang=en (accessed
August 1, 2009).
Campbell, H. 2009. LOHAFEX: If you mean well,
are you allowed to screw up the ocean? Available
online at: http://www.scientificblogging.com/
science_20/blog/lohafex_if_you_mean_well_
are_you_allowed_screw_ocean (accessed
August 1, 2009).
Canadian SOLAS Report. 2003. A research
network for the Canadian Surface Ocean Lower
Atmosphere Study. Progress Report 2001–2003.
April 30, 2003, C-SOLAS Secretariat, Halifax, Nova
Scotia, 54 pp. Available online at: http://as01.ucis.
dal.ca/wag/template/uploads/solas/SOLASreport_
may1.doc.pdf (accessed August 10, 2009).
Carboncorp USA. 1999. Ocean carbon biomass
sequestration. Executive summary. Available online
at: http://web.archive.org/web/20000112024658/
http://rsrch.com/carboncorp/execsummary2.html
Charlson, R.J., J.E. Lovelock, M.O. Andreae, and S.G.
Warren. 1987. Oceanic phytoplankton, atmospheric sulphur, cloud albedo, and climate. Nature
326:655–661.
Chisholm, S.W. 2000. Stirring times in the Southern
Ocean. Nature 407:685-686.
Chisholm, S.W., and F.M.M. Morel, eds. 1991.
What regulates phytoplankton production in
nutrient‑rich areas of the open sea? Limnology and
Oceanography Special Edition 36(8), 493 pp.
Chisholm, S.W., P.G. Falkowski, and J.J. Cullen. 2001.
Oceans – Dis-crediting ocean fertilization. Science
294(5541):309–310.
Chu, S., S. Elliott, M. Maltrud, F. Chai, and F. Chavez.
2003. Studies of the consequences of carbon
sequestration using Los Alamos ocean models.
Paper presented at the Second National Energy
and Technology Laboratory Conference on Carbon
Sequestration. Washington, DC. Available online
at: http://www.netldoe.gov/publications/proceedings/03/carbon-seq/PDFs/043.pdf (accessed
August 6, 2009).
Climate Response Fund. 2009. Available online
at: http://www.oceanstewardship.com/bios/
Margaret%20Leinen.pdf (accessed July 30, 2009).
Climos About Us. 2008. Team. Available online at:
http://www.climos.com/team1.php (accessed
August 3, 2009).
Climos FAQ. 2008. Available online at: http://www.
climos.com/faq.php (accessed July 30, 2009).
Climos Press. 2007. Climos receives first methodology
for ocean iron fertilization from EcoSecurities,
Signs with DNV for Assessment. December 3,
2007. Available online at: http://www.climos.com/
press_detail.php?pid=55 (accessed August 1, 2009).
Coale, K.H., K.S. Johnson, F.P. Chavez, K.O. Buesseler,
R.T. Barber, M.A. Brzezinski, W.P. Cochlan, F.J.
Millero, P.G. Falkowski, J.E. Bauer, and others.
2004. Southern ocean iron enrichment experiment:
Carbon cycling in high- and low-Si waters. Science
304(5669):408–414.
Coale, K.H., K.S. Johnson, S.E. Fitzwater,
R.M. Gordon, S. Tanner, F.P. Chavez, L. Ferioli,
C. Sakamoto, P. Rogers, F. Millero, and others.
1996. A massive phytoplankton bloom induced
by an ecosystem-scale iron fertilization
experiment in the equatorial Pacific Ocean.
Nature 383(6600):495–501.
Consumer Eroski. 2007. El buque “Weatherbird II”
no vertera hierro en aguas de Canarias. Available
online at: http://www.consumer.es/web/medio_
ambiente/2007/12/14/172992.php (accessed
April 24, 2009).
Courtland, R. 2008. Planktos dead in the water.
Nature News. Available online at: http://www.bioedonline.org/news/news.cfm?art=3823 (accessed
August 1, 2009).
Cullen, J.J., and P.W. Boyd. 2008. Predicting and
verifying the intended and unintended consequences of large-scale ocean fertilization. Marine
Ecology Progress Series 364:295–301, doi:10.3354/
meps07551.
Cullen, J.J. 1995. Status of the iron hypothesis after the
open-ocean enrichment experiment. Limnology
and Oceanography 40:1,336–1,343.
de Baar, H.J.W. 1994. Von Liebig’s law of the minimum
and plankton ecology (1899–1991). Progress in
Oceanography 33:347–386
de Baar, H.J.W., P.W. Boyd, K.H. Coale, M.R. Landry,
A. Tsuda, P. Assmy, D.C.E. Bakker, Y. Bozec,
R. Barber, and others. 2005. Synthesis of iron
fertilization experiments: From the Iron Age in
the Age of Enlightenment. Journal of Geophysical
Research 110, C09S16, doi:10.1029/2004JC002601.
Denman, K. 2008. Climate change, ocean processes,
and ocean iron fertilization. Marine Ecology
Progress Series 234:219–225.
Department of Energy. 1999. Enhancement of natural
carbon sequestration in the ocean. Chapter 3.2 in
Carbon Sequestration: Research and Development,
a 1999 report by DOE’s Office of Fossil Energy
and Office of Science, Washington, DC. Available
online at: http://www.fossil.energy.gov/programs/
sequestration/publications/1999_rdreport/
(accessed August 7, 2009).
Doney, S.C., M.R. Abbott, J.J. Cullen, D.M. Karl, and
L. Rothstein. 2004. From genes to ecosystems: The
ocean’s new frontier. Frontiers in Ecology and the
Environment 2:457–466.
Environmental Protection Agency. 2007. Submission
to 30th Meeting of the Scientific Group of the
London Convention. Document LC/SG 30/
INF.28. Planktos, Inc. Large-scale Ocean Iron
Addition Projects.
ETC Group News Release. 2007. Planktos’ commercial
ocean iron fertilization carbon-trading gambit:
Brakes on flakes. Available online at: http://www.
etcgroup.org/en/materials/publications.html?pub_
id=641 (accessed March 25, 2009).
ETC Group News Release. 2009. LOHAFEX Update:
Geo-engineering ship plows on as Environment
Ministry calls for a halt. 2009. January 13, 2009.
Available online at: http://www.etcgroup.org/en/
materials/publications.html?pub_id=712 (accessed
March 25, 2009).
Fuhrman, J.A., and D.G. Capone. 1991. Possible
biogeochemical consequences of ocean
fertilization. Limnology and Oceanography
36(8):1,951–1,959.
Oceanography
September 2009
259
George, R. 2007. Information provided to the House
Select Committee on Energy Independence and
Global Warming. Wednesday, July 17, 2007.
Available online at: http://74.125.45.132/search?q
=cache:n1jTbvaQsEAJ:globalwarming.house.gov/
tools/assets/files/0026.pdf+Russ+George+
Planktos+restoration+2006&cd=1&hl=en&ct=clnk
&gl=us (accessed on May 18, 2009).
Gnanadesikan, A., and I. Marinov. 2008. Export is not
enough: Nutrient cycling and carbon sequestration. Marine Ecology Progress Series 364:289–294,
doi: 10.3354/meps07550.
Gnanadesikan, A., J.L. Sarmiento, and R.D. Slater.
2003. Effects of patchy ocean fertilization on
atmospheric carbon dioxide and biological production. Global Biogeochemical Cycles 17(2), 1050,
doi:10.1029/2002GB001940.
Graeber, C. 2000. Dumping iron: Ecohacker Michael
Markels claims he has a megafix for global
warming: Supercharge the growth of ocean
plankton with vitamin Fe and let a zillion CO2
scrubbers bloom. Wired. Available online at: http://
www.wired.com/wired/archive/8.11/ecohacking_
pr.html (accessed August 1, 2009).
Hardin, G. 1968. The tragedy of the commons.
Science 162:1,243–1,248.
Hart, T. 1934. On the plankton of the Southwest
Atlantic and the Bellingshausen Sea. Discovery
Reports 8:1–268.
Hoffmann, L.J., I. Peeken, K. Lochte, P. Assmy,
and M. Veldhuis. 2006. Different reactions of
Southern Ocean phytoplankton size classes to
iron fertilization. Limnology and Oceanography
51(3):1,217–1,229.
Holland, G. 2001. Workshop on Ocean Fertilization
to Transfer Atmospheric CO2 to the Oceans.
Paper presented at the American Society for
Limnology and Oceanography Ocean Fertilization
Symposium, Washington, DC, April 25, 2001.
Industrial Environment. 2006. California green groups
embrace Planktos ecorestoration. Company overview. November 1, 2006. Available online at: http://
www.accessmylibrary.com/coms2/summary_028632130890_ITM (accessed May 18, 2009).
International Center for Technology Assessment. 2007.
Letter to the Securities Exchange Commission
regarding false and misleading statements made
by Planktos, Corp. June 22, 2007. Available
online at: http://www.icta.org/global/actions.
cfm?page=1&type=369&topic=12 (accessed
April 24, 2009).
Jacquet, S.H.M., N. Savoye, F. Dehairs, V.H. Strass, and
D. Cardinal. 2008. Mesopelagic carbon remineralization during the European iron fertilization
experiment. Global Biogeochemical Cycles 22,
GB1023, doi:10.1029/2006GB002902.
Jin, X., N. Gruber, H. Frenzel, S.C. Doney, and
J.C. McWilliams. 2008. The impact on atmospheric CO2 of iron fertilization induced
changes in the ocean’s biological pump.
Biogeosciences 5(2):385–406.
Johnson, K.S., and D.M. Karl. 2002. Is ocean
fertilization credible and creditable?
Science 296(5567):467–467.
260
Oceanography
Vol.22, No.3
Karl, D.M., and R.M. Letelier. 2008. Nitrogen fixationenhanced carbon sequestration in low nitrate, low
chlorophyll seascapes. Marine Ecology Progress
Series 364:257–268.
Keim, B. 2009. Ocean fertilization works—unless
it doesn’t. Wired Science. March 25, 2009.
Available online at: http://blog.wired.com/
wiredscience/2009/01/oceaniron.html (accessed
August 1, 2009).
Kerry, S. 2008. Planktos announces resignation and
release of CEO and employees. Iron Fertilization
News. March 25, 2009. Available online at: http://
carbonsequestration.blogspot.com/2008/04/
planktos-announces-resignation-and.html
(accessed August 1, 2009).
Kintisch, E. 2009. Controversy in Their Wake,
Geoengineering Experiment in Southern Ocean
to Begin. Science Insider. January 26, 2009.
Available online at: http://blogs.sciencemag.org
/scienceinsider/2009/01/controversy-in.html
(accessed August 1, 2009).
Kohfeld, K.E., C. Le Quere, S.P. Harrison, and R.F.
Anderson. 2005. Role of marine biology in glacialinterglacial CO2 cycles. Science 308(5718):74–78.
Krivit, S.B. 2007. The Wizard of Half Moon Bay: A
New Energy Times special report on Planktos
and D2Fusion. New Energy Times Issue 24,
September 10, 2007. Available online at: http://
newenergytimes.com/v2/news/2007/NET24.
shtml#thewiz (accessed August 1, 2009).
Kunzig, R. 1994. The Iron Man’s revenge. Discover June
1994. Available online at: http://discovermagazine.
com/1994/jun/theironmansreven391 (accessed
August 1, 2009).
Lampitt, R.S., E.P. Achterberg, T.R. Anderson,
J.A Hughes, M.D Iglesias-Rodriguez, B.A. KellyGerreyn, M. Lucas, E.E Popova, R. Sanders,
J.G Shepherd, and others. 2008. Ocean fertilization: A potential means of geoengineering?
Philosophical Transactions of the Royal
Society A 366:3,919–3,945.
Latham, J., P. Rasch, C.C. Chen, L. Kettles, A. Gadian,
A. Gettelman, H. Morrison, K. Bower, and
T. Choularton. 2008. Global temperature stabilization via controlled albedo enhancement of lowlevel maritime clouds. Philosophical Transactions of
the Royal Society A 366:3,969–3,987.
Law, C.S., and the SAGE Team. 2006. Plankton, iron,
and climate. Water and Atmosphere 14(2):21.
Available online at: http://www.niwa.co.nz/newsand-publications/publications/all/wa/14-2/iron
(accessed August 1, 2009).
Law, C.S. 2008. Predicting and monitoring the effects
of large-scale ocean iron fertilization on marine
trace gas emissions. Marine Ecology Progress
Series 364:283–288.
Leinen, M., K. Wilden, and D. Whaley. 2008. A
response to concerns raised about ocean iron fertilization by Greepeace. Available online at: http://
www.climos.com/imo/Climos/Climos_Response_
to_Greenpeace_OIF_May2008_Final.pdf (accessed
August 1, 2009).
London Convention Meeting Report. 2007. Report
of the Twenty-Ninth Consultative Meeting and
the Second Meeting of the Contracting Parties.
December 14, 2007. Available online at: http://
www.imo.org/includes/blastDataOnly.asp/data_
id%3D20797/17.pdf (accessed August 4, 2009).
London Convention Meeting Report. 2008. Report
of the Thirtieth Consultative Meeting and
the Third Meeting of the Contracting Parties.
December 9, 2008. Available online at: http://
www.imo.org/includes/blastDataOnly.asp/data_
id%3D24255/16.pdf (accessed August 4, 2009).
Markels, M. 1995. US Patent 5433173: Method of
improving production of seafood. Available online
at: http://www.patentstorm.us/patents/5433173/
fulltext.html (accessed August 1, 2009).
Markels, M. 1998. Farming the Oceans: An update.
Regulation 21(2):9–10. Available online at:
http://www.cato.org/pubs/regulation/regv21n2/
v21n2.html (accessed August 1, 2009).
Markels, M. 1999. Effects of creativity and entrepreneurship. Cosmos Journal. Available online at:
http://www.cosmos-club.org/web/journals/1999/
markels.html (accessed August 1, 2009).
Markels, M., and R. Barber. 2001. Sequestration
of CO2 by ocean fertilization. Paper presented
at the First National Energy and Technology
Laboratory Conference on Carbon Sequestration,
Washington, DC.
Martin, J.H. 1990a. A new iron age, or a ferric fantasy.
JGOFS Newsletter 1(4):5,11.
Martin, J.H. 1990b. Glacial-interglacial CO2 exchange:
The iron hypothesis. Paleoceanography 5:1–13.
Martin, J.H., K.H. Coale, K.S. Johnson, S.E. Fitzwater,
R.M. Gordon, S.J. Tanner, C.N. Hunter, V.A. Elrod,
J.L. Nowicki, T.L. Coley, and others. 1994. Testing
the iron hypothesis in ecosystems of the equatorial
Pacific Ocean. Nature 371(6493):123–129.
Martin, J. H., R.M. Gordon, S. Fitzwater, and
W.W. Broenkow. 1989. Vertex: Phytoplankton iron
studies in the Gulf of Alaska. Deep-Sea Research
Part A 36(5):649–680.
Martin, J.H., R.M. Gordon, and S.E. Fitzwater.
1990. Iron in Antarctic waters. Nature
345(6271):156–158.
McKie, R. 2003. Bid to reduce greenhouse gases ‘is
folly.’ The Observer. January 12, 2003. Available
online at: http://www.guardian.co.uk/uk/2003/
jan/12/research.science (accessed August 1, 2009).
Monastersky, R. 1994. Pumping iron: Too weak to slow
warming. Science News 145(10):148.
Morgan, J. 2009. Ocean climate fix remains afloat.
BBC News: Science and Environment. January 29,
2009. Available online at: http://news.bbc.
co.uk/2/hi/science/nature/7856144.stm (accessed
August 1, 2009).
Murray, J. 2008. Climos aiming for first “ocean fert”
project in late 2009. BusinessGreen. September
23, 2008. Available online at: http://www.
businessgreen.com/business-green/news/2226690/
climos-aiming-first-ocean-fert (accessed
August 1, 2009).
NIO Press Release. 2009. LOHAFEX provides
surprising insights on plankton ecology that
dampen hopes of using the Southern Ocean
to sequester atmospheric CO2. National
Institute of Oceanography. Available online
at: http://www.nio.org/projects/narvekar/
Lohafex_press_release_23_03_09.pdf (accessed
August 1, 2009).
Owens, N.J.P. 2009. Review of: LOHAFEX Iron
Enrichment Experiment, Southwest Atlantic.
Available online at: http://www.bmbf.de/_media/
press/BAS_zu_LOHAFEX.pdf (accessed
August 1, 2009).
Parks, N. 2008. Fertilizing the seas for climate mitigation: Promising strategy or sheer folly? American
Institute of Biological Sciences Washington
Watch. Available online at: http://www.aibs.org/
washington-watch/washington_watch_2008_
02.html (accessed August 1, 2009).
Paytan, A., M. Kastner, and F.P. Chavez. 1996. Glacial
to interglacial fluctuations in productivity in the
equatorial Pacific as indicated by marine barite.
Science 274(5291):1,355–1,357.
Planktos Green Tags. 2002. Originally from www.
planktos.com/greentags.htm. Available online
at: http://web.archive.org/web/20020612160156/
www.planktos.com/greentags.htm (accessed
August 1, 2009).
Plotkin, H. 2002. Ocean Rescue: Planktos Foundation
hopes to reduce global warming by fertilizing the
seas. San Francisco Chronicle: SF Gate. May 22,
2002. Available online at: http://www.sfgate.com/
cgi-bin/article.cgi?file=/gate/archive/2002/05/22/
plnktos.DTL (accessed August 2, 2009).
Powell, H. 2007–2008. Six-part series. Part 1:
Fertilizing the oceans with iron. Part 2: Will
ocean iron fertilization work? Part 3: What are
the possible side effects? Part 4: Lessons from
nature, models, and the past. Part 5: Dumping iron
and trading carbon. Part 6: Proposals emerge to
transfer excess carbon into the ocean. Oceanus.
November 2007–January 2008. Available online
at: http://www.whoi.edu/oceanus/viewArticle.
do?id=34167 (accessed August 2, 2009).
Rasch, P.J., P.J. Crutzen, and D.B. Coleman. 2008.
Exploring the geoengineering of climate using
stratospheric aerosols: The role of particle
size. Geophysical Research Letters 35, L02809,
doi:10.1029/2007GL032179.
Rees, F.P., P.D. Nightingale, N.J.P. Owens, and the
PML FeeP Team. 2007. FeeP-An in situ PO43- and
Fe2+ addition experiment to waters of the subtropical north-east Atlantic. Geophysical Research
Abstracts 9:A01440.
Report of the First Meeting of the Intersessional
Legal and Related Issues Working Group on
Ocean Fertilization. 2009. February 11–13, 2009,
28 pp. Available online at: http://www.imo.org/
Safety/mainframe.asp?topic_id=1777 (accessed
August 7, 2009).
Report of the First Meeting of the Intersessional
Technical Working Group on Ocean Fertilization.
2009. Available online at: www.imo.org/includes/
blastData.asp/doc_id=11172/5.pdf (accessed
August 1, 2009).
Rice, M.L. 2003. GreenSea’s interest in fertilizing sea
with iron. Nature 421(6925):786–786.
Riddell, L. 2008. Climos aims to profit from hungry
plankton. San Francisco Business Times. April 25,
2008. Available online at: http://www.bizjournals.
com/sanfrancisco/stories/2008/04/28/story17.html
(accessed August 2, 2009).
San Francisco Business Times. 2007. Planktos starts
to sink. December 19, 2007. Available online
at: http://www.bizjournals.com/sanfrancisco/
stories/2007/12/17/daily28.html (accessed
August 3, 2009).
Sarmiento, J.L., and J.C. Orr. 1991. Three-dimensional
simulations of the impact of Southern Ocean
nutrient depletion on atmospheric CO2. Limnology
and Oceanography 36:1,928–1,950.
Schiermeier, Q. 2003. Climate change: The oresmen.
Nature 421(6919):109–110.
Scientific Group of the London Convention Meeting
Report. 2007. Report of the Thirtieth Meeting of
the Scientific Group of the London Convention and
First Meeting of the Scientific Group of the London
Protocol. July 25, 2007. Available online at: http://
www.imo.org/includes/blastDataOnly.asp/data_
id%3D19765/14.pdf (accessed August 4, 2009).
Scientific Group of the London Convention Meeting
Report. 2008. Report of the Thirty-First Meeting
of the Scientific Group of the London Convention
and Second Meeting of the Scientific Group of the
London Protocol. July 8, 2008. Available online
at: http://www.imo.org/includes/blastDataOnly.
asp/data_id%3D22812/16.pdf (accessed
August 4, 2009).
Sea Shepherd News. 2008. Sea Shepherd defeats the
Planktos ocean dumping scheme. Sea Shepherd
News, February 14, 2008. Available online at:
http://www.seashepherd.org/news-and-media/
news-080214-1.html (accessed August 2, 2009).
South Bay. 2007. Sea Shepherd to intercept/
block Planktos ship. South Bay, July 28, 2007.
Available online at: http://www.indybay.org/
newsitems/2007/07/28/18437803.php (accessed
August 2, 2009).
Smetacek, V. 2001. EisenEx: International team
conducts iron experiment in Southern Ocean.
US JGOFS News 11(1):11–14.
Smetacek, V., and S.W.A. Naqvi. 2008. The next
generation of iron fertilization experiments in the
Southern Ocean. Philosophical Transactions of the
Royal Society A 366:3,947–3,967.
Sullivan, K. 2007. World Wildlife Fund warns against
iron dumping experiment near the Galapagos
Islands. WWF Press Release. Available online
at: http://www.worldwildlife.org/who/media/
press/2007/WWFPresitem973.html (accessed
August 1, 2009).
Suplee, C. 1996. Iron supplement for a cooler world.
The Washington Post, October 14, 1996.
Szabo, M. 2009. New science could help solve climate
crisis. Reuters: Environment News. Available
online at: http://www.reuters.com/article/
environmentNews/idUSTRE50R08V20090128
(accessed August 1, 2009).
Thompson, K. 2008. Carbon discredit. Popular
Science. Available online at: http://www.popsci.
com/environment/article/2008-07/carbondiscredit?page=4 (accessed April 24, 2009).
Toggweiler, J.R., J.L. Russell, and S.R. Carson.
2006. Midlatitude westerlies, atmospheric
CO2, and climate change during the ice ages.
Paleoceanography 21, doi:10.1029/2005PA001154.
Tsuda, A., H. Kiyosawa, A. Kuwata, M. Mochizuki,
N. Shiga, H. Saito, S. Chiba, K. Imai, J. Nishioka,
and T. Ono. 2005. Responses of diatoms to ironenrichment (SEEDS) in the western subarctic
Pacific, temporal and spatial comparisons. Progress
in Oceanography 64(2–4):189–205.
Tsuda, A., S. Takeda, H. Saito, J. Nishioka, I. Kudo,
Y. Nojiri, K. Suzuki, M. Uematsu, M.L. Wells,
D. Tsumune, and others. 2007. Evidence for the
grazing hypothesis: Grazing reduces phytoplankton
responses of the HNLC ecosystem to iron enrichment in the western subarctic pacific (SEEDS II).
Journal of Oceanography 63(6):983–994.
UN CBD. 2008. Biodiversity and Climate Change:
Ocean Fertilization. Decision IX.16 UN
Convention on Biological Diversity. Adopted
May 31, 2008.
Williams, W. 2003. Iron fertilization. Living on Earth.
Available online at: http://www.loe.org/series/
iron_fertilization/ (accessed August 1, 2009).
Winckler, G., R.F. Anderson, M.Q. Fleisher, D. Mcgee,
and N. Mahowald. 2008. Covariant glacialinterglacial dust fluxes in the equatorial Pacific and
Antarctica. Science 320(5872):93–96.
Woese, C.R., and N. Goldenfeld. 2009. How the
microbial world saved evolution from the Scylla
of molecular biology and the charybdis of the
modern synthesis. Microbiology and Molecular
Biology Reviews 73:14–21.
World Wildlife Fund International. 2009. WWF
opposes precarious ocean fertilization project.
Available online at: http://www.panda.org/wwf_
news/news/?155301/WWF-opposes-precariousocean-fertilization-project (accessed July 27, 2009).
Zahariev, K., J.R. Christian, and K.L. Denman.
2008. Preindustrial, historical, and fertilization
simulations using a global ocean carbon model
with new parameterizations of iron limitation, calcification, and N2 fixation. Progress in
Oceanography 77(1):56–82.
Oceanography
September 2009
261
Download